Está en la página 1de 495

Vertigo

Springer-Verlag London Ltd.

Thomas Brandt

Vertigo
Its Multisensory Syndromes

2nd Edition

Springer

Professor Thomas Brandt, FRCP


Neurologische Klinik, Klinikum Grohadern, Ludwig-Maximillians-Universitt,
Marchioninistrae 15,81377 Munich, Germany

ISBN 978-0-387-40500-1
British Library Cataloguing in Publication Data
Brandt, Thomas
Vertigo: its multisensory syndromes. - 2nd ed.
1. Vertigo 2. Diagnosis, Differential.
I. Title
616.8'41
ISBN 978-0-387-40500-1
ISBN 978-1-4757-3801-8 (eBook)
DOI 10.1007/978-1-4757-3801-8
Library of Congress Cataloging-in-Publication Data
A catalog record for this book is available from the Library of Congress
Apart from any fair dealing for the purposes of research or private study, or criticism or review, as
permitted under the Copyright, Designs and Patents Act 1988, this publication may only be
reproduced, stored or transmitted, in any form or by any means, with the prior permission in writing
of the publishers, or in the case of reprographic reproduction in accordance with the terms of licences
issued by the Copyright Licensing Agency. Enquiries concerning reproduction outside those terms
should be sent to the publishers.
Springer-Verlag London 2003
Ursprnglich erschienen bei Springer-Verlag London Limited 2003

1st edition published in 1991


The use of registered names, trademarks, etc. in this publication does not imply, even in the absence
of a specific statement, that such names are exempt from the relevant laws and regulations and
therefore free for general use.
Product liability: The publisher can give no guarantee for information about drug dosage and
application thereof contained in this book. In every individual case the respective user must check its
accuracy by consulting other pharmaceuticalliterature.
Typeset by Columns Design Ud, Reading
28/3830-543210 Printed on acid-free paper SPIN 10943052

Preface to the Second Edition

This monograph has been written for clinicians who are involved in the management of the
dizzy patient and for scientists with a particular interest in the multi-sensorimotor mechanisms that subserve spatial orientation, motion perception, and ocular motor and postural control. Special emphasis has been put on making the correct diagnosis, and detailed
recommendations have been given for specific treatments.
The second edition has resulted in an almost completely new book due to the dramatic
expansion in the 1990s of our understanding of vestibular function and dis orders. A few relevant examples include the novel concept of canalolithiasis, as opposed to cupulolithiasis, both
of which are established causes of typical posterior and horizontal canal benign paroxysmal
positioning vertigo; familial episodic ataxia land II have been identified as inherited channelopathies; otolithic syndromes were recognized as a variety separate from semicircular
canal syndromes; several new central vestibular syndromes have been described, localized,
and attributed to vestibular pathways and centres; a new classification based on the three
major planes of action of the vestibulo-ocular reflex is available for central vestibular syndromes; and the mystery of the location and function of the multisensory vestibular cortex is
slowly being unravelled.
This book differs from other clinical textbooks in that it is not divided into two parts:
anatomy and physiology, on the one hand, and disorders, on the other. Introductory chapters
on several aspects of vestibular syndromes, their diagnosis, and their management are followed by sections and chapters that focus on the description of specific dis orders. Anatomy
and physiology are discussed only when relevant for the understanding of the mechanism.
Although there are many experts in the field who know better than I particular diseases of
their interest, I nevertheless ventured on the writing of this interdisciplinary book alone for
two reasons: first, to make the reader's usage easier by a uniform presentation and second, to
improve my own competence in treating the dizzy patient by studying the research of others.
The central focus of the book is on the patient who because of complaints that are typical of
different disorders is frequently shuttled between neurologists, otolaryngologists, internists,
and psychiatrists.

Preface to the First Edition

Vertigo consists of a variety of syndromes which are surprisingly easy to diagnose and can, in
most cases, be treated effectively. However treatment requires an interdisciplinary approach to
the patient which is unusual for clinicians who have usually been trained to specialise in a
particular area. Sensorimotor physiology is the key to an understanding of the pathogenesis
of vertigo; careful history-taking and otoneurological examination are the key to diagnosis.
The book is organised in sections covering the major sub divisions of vertigo, including
peripherallabyrinthine disorders (Meniere's disease, vestibular neuritis, perilymph fistulas),
central vestibular dis orders (vestibular epilepsy, downbeat/upbeat nystagmus), positional,
vascular, traumatic and familial vertigo, vertigo in childhood and vertigo related to drugs.
Sections are further subdivided into chapters covering particular aspects, for example the
chapter on migraine and vertigo in the section on vascular vertigo. There is a full description
ofthe clinical features and diagnostic procedures for each disease (with summarising tables),
and special emphasis is placed on the relationship between management and the underlying
pathological mechanisms.
Most diseases are referred to in several different sections in order to facilitate the differential diagnosis of conditions with similar signs and symptoms. The section on vertigo arising
from multisensory interaction covers non-vestibular syndromes such as visual vertigo and
cervical vertigo and, more importantly, the psychogenic vertigo syndromes; the latter are the
third commonest cause of vertigo in patients seen by neurologists.
This book will contribute to an improvement in diagnosis and management in patients suffering from vertigo and disequilibrium. A further dem an ding goal of this book is to establish
a platform from which physiologists and clinicians may launch cooperative research concerning the intriguing mechanisms of spatial orientation, oculomotor and postural control and
ultimately to aid patients with vertigo.

Acknowledgements

I am indebted to many people for helping with this second edition. I want first to thank Judy
Benson, who not only conscientiously undertook the language editing of the manuscript but
also as an attentive reader, unburdened by a medical background, gave valuable impulses for
resolving ambiguities and unclarities. Michael Strupp, an experienced colleague in our
Dizziness Unit, made himself indispensable. He critically read the entire manuscript, made
important suggestions for improvement, and drew my attention to missing details and relevant references. Thanks are also due to MicheIe Seiche, who carefully cross-checked citations
in the text and typed and proofed the references.
I wish to express my special thanks to my colleagues in the Dizziness Unit for the stimulating daily discussions on which a large part of our clinical experience is based, in particular
Marianne Dieterich, who heads the clinical research group on ocular motor and vestibular
disorders.
I am grateful for the constructive cooperation I enjoyed with Springer-Verlag London, in
particular Christopher GreenweIl. I also sincerely thank the rest of the staff of Springer-Verlag
for their efforts to meet our pressing deadlines. Last, but certainly not least, I want to express
my gratitude to my family for their understanding during the ordeal, above all to my wife
Birgit, who knowing how important this project was to me, gave it her full support, deferring
her own interests and wishes for the sake of its completion.

Contents

Glossary .............................................................. xxiii

Seetion A Vertigo: symptoms, syndromes, dis orders ..................


1 Introduction

The "vestibular" vertigo syndromes ................................


Signs and symptoms ...........................................
The mismatch concept .........................................
The vestibulo-ocular reflex (VOR) .................................
Neuronal network of the VOR. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
VOR mediation of perception and postural adjustments ............
Vestibulocollic reflex ...........................................
Vestibulospinal reflexes ...........................................
Vestibular falls ..................................................
Peripheral vestibular falls .......................................
Vestibular neuritis: contraversive rotational vertigo with
ipsiversive falls ............................................
Benign paroxysmal positioning vertigo (BPPV): forward falls
produced by canalolithiasis of the posterior semicircular canal "
Meniere's drop attacks (Tumarkin's otolithic crisis) ..............
Otolith Tullio phenomenon: contraversive ocular tilt
reaction (OTR) and fall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Bilateral vestibulopathy with predominant forward and
backward falls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Central vestibular falls .........................................
Vestibular epilepsy with contraversive vertigo and falls ...........
Thalamic astasia with contraversive or ipsiversive falls? ...........
Ocular tilt reaction: ipsiversive in caudal, contraversive in
upper brainstem lesions ....................................
Lateropulsion in Wallenberg's syndrome: ipsiversive falls and
adjustments of perceived vertical ............................
Downbeat nystagmus syndrome with backward falls .............
Vestibular autonomie regulation ...................................
Neuroanatomie substrates ......................................
References ........................................................

1
3

3
4
4
5
7

9
10
10
13
14
14
15
15
15
15
15
15
16
16
16
16
16
18
19

xii

Contents

Approaching the patient

23

Dizziness and light-headedness ..................................


Attacks of (rotatory) vertigo, episodic vertigo .....................
Sustained (rotatory) vertigo .....................................
Positional/positioning vertigo ...................................
Oscillopsia ....................................................
Vertigo associated with auditory dysfunction ......................
Dizziness or to-and-fro vertigo and postural imbalance .............
Semicircular canal vertigo and mixed canal-otolith vertigo ............
Otolithic vertigo .................................................
Paroxysmal vertigo ..............................................
Neuro-ophthalmological and otoneurological evaluation ..............
References ........................................................

23
23
24
26
26
27
27
28
29
29
34
47

3 Management of the dizzy patient ....................................

49

Antivertiginous and antiemetic drugs ..............................


Surgical treatment ...............................................
Vestibular exercises and physical therapy for vestibular rehabilitation ..
Quantitative effects of balance training on postural sway in
normal subjects .............................................
Balance training in vestibular disorders ...........................
Plasticity of the vestibular system: central compensation and
sensory substitution for vestibular deficits ......................
Terms and definitions of plasticity and central compensation ........
Vestibular compensation and its multiple mechanisms ..............
Transmitters of the vestibulo-ocular reflex and drug-modulated
compensation ...............................................
Substitution of vestibular function ...............................
References ........................................................

49
51
52

Section B
4

52
53
55
55
56
58
60
61

Vestibular nerve and labyrinthine dis orders ...............

65

Vestibular neuritis .................................................


The clinical syndrome ............................................
Vertigo and posture ............................................
Eye movements ................................................
Caloric testing ................................................
MR imaging ..................................................
Natural course ................................................
High-frequency defect ofVOR in permanent peripheral
vestibular lesion .............................................
Differential diagnosis ..........................................
Aetiology and pathomechanism ...................................
Pathomechanism ..............................................
Vestibular neuritis - a partial unilateral vestibular loss .............
Viral or vascular aetiology? .....................................
Historical discussion .........................................
Arguments for viral aetiology .................................

67
67
68
69
69
70
71
71
72
73
73
73
75
75
75

Contents

xiii

Site of the lesion


Management ....................................................
References ........................................................

76
76
79

5 Meniere's disease ..................................................

83

The clinical syndrome ............................................


Attacks .......................................................
Auditory symptoms and signs in the vertigo-free interval ...........
Vestibular function in the vertigo-free interval .....................
Imaging ......................................................
Differential diagnosis ..........................................
Natural course. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Aetiology and pathomechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Endolymphatic hydrops ........................................
Aetiology .....................................................
Delayed endolymphatic hydrops ...............................
Vascular hypothesis ..........................................
Psychosomatic hypothesis ....................................
Pathophysiology of attacks and progressive dysfunction . . . . . . . . . . . . .
Management ....................................................
Attacks .......................................................
Attack-free interval . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Intratympanic gentamicin therapy ...............................
Surgical treatments: nondestructive or destructive . . . . . . . . . . . . . . . . . .
Non-destructive .............................................
Destructive .................................................
Pragmatic therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Vestibular drop attacks (Tumarkin's otolithic crisis) ..................
References ........................................................

83
83
84
85
85
85
85
86
86
87
88
88
89
89
90
90
91
91
92
92
93
93
94
95

Perilymph fistulas (PLF) ............................................

99

The clinical syndromes ...........................................


Semicircular canal type of PLF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Otolith type of PLF ............................................
How maya perilymph fistula be identified? ........................
Pressure fistula tests ........................................ "
Vascular fistula tests. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Imaging techniques (CT, MRI) ................................
Electronystagmography ......................................
Hearing tests. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Exploratory tympanotomy ...................................
Other proposed tests .........................................
Differential diagnosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Aetiology and pathomechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Experimental perilymph/endolymph fistulas and
endolymphatic hydrops .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Management ....................................................

99
100
100
101
101
101
102
102
102
102
102
102
102

104
105

xiv

Contents

Conservative treatment .........................................


Surgical treatment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Tullio phenomenon ..............................................
Experimental history . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Clinical history ... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Clinical types of Tullio phenomena . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Otolith Tullio phenomenon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Why does otolith Tullio phenomenon manifest with
paroxysmal ocular tiIt re action (OTR)? .......................
Vestibulospinal reflexes tested as part of the Tullio phenomenon .....
Management ..................................................
Fistula of the anterior semicircular canal ............................
References ........................................................

105
106
106
106
107
107
107
108
108
111
112
113

7 Peripheral vestibular paroxysmia (disabling positional vertigo) . . . . . . . . .. 117


The clinical syndrome ............................................
Case reports . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Vertigo .....................................................
Auditory symptoms and tests. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Other associated symptoms ...................................
Electronystagmography ......................................
Subjective visual vertical . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Posturography. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Differential diagnosis ..........................................
Aetiology and pathomechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Management ....................................................
Uncertainties in the diagnosis and treatment of vestibular
paroxysmia .................................................
AIternating episodes of vestibular nerve paroxysmia and failure . . . . . . ..
References ........................................................

117
118
120
120
120
121
121
121
121
122
122
123
124
125

8 Bilateral vestibulopathy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 127


The clinical syndrome ............................................
Diagnosis .....................................................
Associated symptoms and differential diagnosis . . . . . . . . . . . . . . . . . . ..
Aetiologies and pathomechanisms .................................
Idiopathic BVF ..............................................
Spatial orientation: vestibulo-ocular and vestibulospinal reflexes .....
Management ..................... , ..............................
References ........................................................

127
128
129
129
132
135
137
139

9 Miscellaneous vestibular nerve and labyrinthine dis orders ............. 143


Imaging of the labyrinth and vestibular nerve . . . . . . . . . . . . . . . . . . . . . . ..
Congenital causes ................................................
Infectious causes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Herpes zoster oticus .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Acute otitis media. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..

143
143
145
146
149

Contents

Specific infections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Cholesteatoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Autoimmune inner ear disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Cogan's syndrome . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
How to monitor activity in Cogan's syndrome ...................
Tumours ... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
References ........................................................

xv

151
151
151
154
155
155
165

Section C Central vestibular disorders ............................... 167


10

Vestibular dis orders in (frontal) roll plane ............................ 175


The clinieal syndrome ............................................
Topographie diagnostie rules ....................................
Ocular tilt reaction (OTR) .......................................
Mechanism of OTR ..........................................
OTR and perceived tilt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Two types of OTR: the medullary "ascending" VOR-OTR and the
mesencephalie "descending" integrator-OTR . . . . . . . . . . . . . . . . . ..
Aetiology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Natural course and management. . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Skew deviation (skew-torsion sign) ...............................
Skew torsion: a vestibular brainstem sign of topographie
diagnostic value ...........................................
Different types of skew deviation ..............................
Alternating skew deviation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Natural course. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Perceived vertieal (subjective visual vertical) . . . . . . . . . . . . . . . . . . . . . ..
Historieal reports on SVV tilts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
SVV tilt - a vestibular sign? ...................................
svv tilt versus room tilt illusion ...............................
SVV tilts in central vestibular versus peripheral ocular
motor lesions .............................................
Thalamic and cortical astasia associated with SVV tilts ...........
Torsional nystagmus. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Three-dimensional modelling of statie vestibulo-ocular
brainstem syndromes ..........................................
References .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..

11

176
178
179
180
181
181
184
185
185
186
186
187
188
189
189
189
190
192
192
193
194
195

Vestibular dis orders in (sagittal) pitch plane. . . . . . . . . . . . . . . . . . . . . . . . . .. 199


Downbeat nystagmus (vestibular downbeat syndrome) . . . . . . . . . . . . . . ..
The clinical syndrome ..........................................
Nystagmus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Oscillopsia and impaired motion perception ....................
Postural imbalance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Aetiology and pathomechanism ..................................
Pathomechanism and site ofthe lesions .........................
Aetiology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..

199
199
200
200
201
201
201
203

xvi

Contents

Management ..................................................
Upbeat nystagmus (vestibular upbeat syndrome) .....................
The clinieal syndrome ..........................................
Nystagmus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Oscillopsia, motion perception, and spatial orientation . . . . . . . . . . ..
Postural imbalance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Aetiology and pathomechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Pathomechanism and site of the lesions . . . . . . . . . . . . . . . . . . . . . . . ..
Aetiology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Management ..................................................
References ........................................................

204
205
206
206
206
207
207
207
208
209
211

12 Vestibular disorders in (horizontal) yaw plane ......................... 215


Horizontal nystagmus as a sign of vestibular tone imbalance
in the yaw plane ............................................... 215
Combined VOR dysfunction in more than one plane of action .......... 217
References ........................................................ 217
13

Vestibular cortex: its locations, functions, and dis orders ................ 219
Multiple vestibular cortex areas ....................................
No primary vestibular cortex ....................................
The parieto-insular vestibular cortex (PIVC) . . . . . . . . . . . . . . . . . . . . . ..
Multimodal sensorimotor vestibular cortex function and dysfunction . ..
Spatial hemineglect, a cortical vestibular syndrome? ................
Paroxysmal room-tilt illusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Self-motion perception: the mechanism of reciprocal inhibitory
visual-vestibular interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
References ........................................................

219
219
220
221
224
224
225
230

14 Vestibular epilepsy ................................................ 233


The vestibular seizure ............................................
Rotatory seizure ("volvular epilepsy") ............................
Differential diagnosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Management ..................................................
Epileptic nystagmus ..............................................
Vestibular versus visual (optokinetie) seizures .......................
"Vestibulogenic epilepsy" .........................................
References ........................................................
15

234
234
234
235
235
237
237
238

Miscellaneous central vestibular dis orders ............................ 241


Central brainstem/cerebellar lesions mimicking vestibular neuritis
or peripheral vestibular failure ....................................
Paroxysmal central vertigo ........................................
Central vestibular falls without vertigo . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Central vestibular syndromes in multiple sclerosis . . . . . . . . . . . . . . . . . . ..
Vestibular syndromes and brain tumours . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Metabolie disorders of the vestibular system .. . . . . . . . . . . . . . . . . . . . . . ..
References ........................................................

241
242
243
244
244
245
245

Contents

Seetion D
16

xvii

Positional and positioning vertigo ......................... 247

Benign paroxysmal positioning vertigo ............................... 251


The clinical syndrome ............................................
Positioning nystagmus .........................................
Vertigo and posture ..............................................
Natural course . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Differential diagnosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Pathomechanism and aetiology .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Pathomechanism ...............................................
Peripheral or central vestibular dysfunction? ....................
The traditional view of cupulolithiasis . . . . . . . . . . . . . . . . . . . . . . . . ..
Arguments for canalolithiasis .................................
Unilateral mimicking bilateral BPPV ...........................
Aetiology .....................................................
Management ....................................................
Positional exercises and liberatory manreuvres . . . . . . . . . . . . . . . . . . . ..
Surgical procedures ............................................
Singular neurectomy .........................................
Plugging of the posterior semicircular canal . . . . . . . . . . . . . . . . . . . ..
Horizontal semicircular canal BPPV (h-BPPV) .......................
The clinical syndrome ..........................................
Atypical h-BPPV with apogeotropic positional nystagmus .........
Natural course. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Aetiology and pathomechanism ..................................
Transition of canalolithiasis to cupulolithiasis ...................
Reversible ipsilateral caloric hypoexcitability .. . . . . . . . . . . . . . . . . ..
Management ..................................................
Anterior semicircular canal BPPV (a-BPPV) .........................
References ........................................................

17

Positional nystagmus/vertigo with specific gravity differential


between cupula and endolymph (buoyancy hypothesis ) ................ 285
Positional alcohol vertigo/nystagmus (PAN) .........................
Positional "heavy water" nystagmus ................................
Positional glycerol nystagmus. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Positional nystagmus with macroglobulinaemia (Waldenstrm's
disease) ......................................................
References ........................................................

18

252
253
254
256
256
257
257
257
257
259
261
264
265
265
269
269
269
269
270
270
271
271
274
275
278
279
280

286
287
287
288
288

Central positional vertigo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 291


Positional downbeating nystagmus .................................
Central positional nystagmus ......................................
Central paroxysmal positional/positioning vertigo and
paroxysmal positioning vomiting ................................
Transient vertebrobasilar ischaemia ................................
Rotational vertebral artery occlusion .............................

291
292
293
296
296

xviii

Contents

Head (neck)-extension vertigo ... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 297


Bending-over vertigo ............................................. 298
References ........................................................ 298

Seetion E Vascular vertigo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 301


19 Stroke and vertigo ................................................. 307
Strokes causing peripheral and central vestibular disorders ............
Anterior inferior cerebellar artery and the internal auditory artery .....
Vertebral artery and posterior inferior cerebellar artery ... . . . . . . . . . . ..
Wallenberg's syndrome .........................................
Basilar artery and paramedian pontine and mesencephalic arteries .....
Vestibular syndromes in roll plane .............................
Vestibular syndromes in pitch plane ............................
Thalamic infarctions .............................................
Cortical infarctions ...............................................
Cortical rotational vertigo .......................................
References ........................................................
20

Migraine and vertigo ............................................... 325


Migraine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
The clinical syndrome . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Aetiology and pathomechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Management ..................................................
Basilar migraine (BM) and "vestibular migraine" . . . . . . . . . . . . . . . . . . . ..
The clinical syndrome ..........................................
Diagnosis of BM with episodic vertigo ("vestibular migraine") . . . ..
Pathomechanisms of vertigo, motion sickness and ocular ...........
motor deficits
Origin of vertigo in migraine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Motion sickness-like symptoms. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Ocular motor deficits in the symptom-free interval
indicate permanent brainstem or cerebellar dysfunction ........
Benign paroxysmal vertigo in childhood ..........................
Benign paroxysmal torticollis in infancy ..........................
Benign recurrent vertigo .......................................
Dizziness and vertigo as facultative symptoms in migraine
apart from BM ............................................ ,.
Association of migraine with other vertigo disorders? . . . . . . . . . . . . . ..
References ........................................................

21

307
308
309
309
312
312
314
314
315
318
322

326
326
326
327
329
329
332
333
333
334
335
335
336
337
337
337
338

Hyperviscosity syndrome and vertigo ................................ 341


The clinical syndrome ............................................
Aetiology and pathomechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Management ....................................................
References ........................................................

341
341
341
342

Section F Traumatic vertigo ......................................... 343


22

Head and neck injury . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 347


Traumatic otolith vertigo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 349
References ........................................................ 349

23

Vertigo due to barotrauma .......................................... 351


Alternobaric vertigo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Blast injury . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Decompression sickness ..........................................
Management ..................................................
Round and oval window fistula caused by barotrauma ................
References ........................................................

24

351
352
352
352
353
354

Iatrogenic vestibular disorders ...................................... 355


Intratympanic gentamicin in Meniere's disease: desired and
undesired effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Quinine: reversible and irreversible side effects ......................
Vertebral artery dissection due to chiropractic neck manipulation . . . . ..
Surgically induced vestibular dysfunction ...........................
Iatrogenic benign paroxysmal positioning vertigo ....................
Vestibular loss associated with chronic noise-induced hearing loss . . . . ..
References ........................................................

355
356
356
356
357
357
358

Section G Hereditary vestibular disorders and vertigo in childhood ... 361


25

Familial periodic ataxia/vertigo (episodic ataxia) ...................... 365


The clinical syndromes ...........................................
Episodic ataxia associated with "interictal" myokymia (type EA-1) ....
Episodic ataxia associated with "interictal" nystagmus (type EA-2) ....
Differential diagnoses ..........................................
Aetiology and pathomechanism ....................................
Episodic ataxia type 1, a potassium channelopathy ..................
Episodic ataxia type-2, a cerebral calcium channelopathy . . . . . . . . . . ..
Effects of acetazolamide and the pathomechanism of EA ............
Management ....................................................
References ........................................................

366
366
367
369
370
370
370
372
372
373

26 Vertigo in childhood ............................................... 375


Benign paroxysmal vertigo of childhood and basilar migraine .........
Motion sickness .................................................
Vestibular neuritis ...............................................
Meniere's disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Perilymph fistulas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Unilateral or bilateralloss of vestibular function .....................
Hereditary dis orders causing peripheral vestibular failure .............
Central vestibular syndromes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
References ........................................................

376
376
376
377
377
377
378
379
379

xx

Contents

Section H
27

Vertigo, dizziness, and falls in the elderly

Vertigo, dizziness, and falls in the elderly ............................. 385


Physiological ageing of the vestibular system ........................
Age-related changes in eye movements and vestibulo-ocular reflexes..
Age-related changes in postural sway and balance ..................
Cautious senile gait and "highest-Ievel gait dis orders" . . . . . . . . . . . . . . . ..
Falls in the elderly . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Dizziness in the elderly ...........................................
References ........................................................

Section I
28

383

385
385
386
386
388
389
391

Drugs and vertigo ......................................... 393

Drugs and vertigo ................................................. 395


Ototoxic agents ..................................................
Cerebellar intoxication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Drugs and eye movements ........................................
References ........................................................

395
397
400
402

Section J Non-vestibular (sensory) vertigo syndromes ................. 405


29

Visual vertigo: visual control of motion and balance ................... 409


Circularvection and linearvection: optokinetically induced
perception of self-motion .........................................
Psychophysics of circularvection .................................
Visual-vestibular interaction: functional significance of
visual and vestibular cortices ..................................
Rollvection-tilt: optokinetic graviceptive mismatch .................
Visual pseudo-Coriolis effect and pseudo-Purkinje effect ..............
Optokinetic motion sickness ......................................
Physiological height vertigo and postural balance ....................
Physical prevention of physiological height vertigo .................
Licence for workers at heights? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
The "visual cliff" phenomenon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Vision and posture ...............................................
Moving visual scenes ...........................................
Visual acuity ..................................................
Near vision and eye-object distance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Visual contral of fore-aft versus lateral body sway ..................
Visual stabilisation in the dark . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Flicker illumination ............................................
Visual field ....................................................
Eye movements, oculomotor dis orders, and postural balance. . . . . . . ..
Nystagmus with oscillopsia impairs balance .......................
Extraocular muscle paresis impairs locomotion and balance .........
Oscillopsia ......................................................
Oscillopsia is smaller than retinal image slip: deficient
vestibulo-ocular reflex .......................................

409
411
413
414
416
417
418
422
422
422
423
424
424
425
426
426
426
427
428
428
429
430
431

Contents

Acquired ocular oscillations with oscillopsia . . . . . . . . . . . . . . . . . . . . . ..


Physiologieal impairment of motion perception with moving eyes .. ..
Normal (physiologieal) inhibitory interactions between
self-motion and object-motion perception . . . . . . . . . . . . . . . . . . . . . ..
Pathologieal (adaptive?) binocular impairment of motion
perception caused by monocular external eye muscle paresis . . . . . ..
Oscillopsia and motion perception in congenital nystagmus .........
References ........................................................
30

xxi

432
433
435
435
435
436

Somatosensory vertigo ............................................. 441


Cervical vertigo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Functional significance of neck afferents and neck reflexes. . . . . . . . . ..
Spatial orientation ...........................................
Neck reflexes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Cervico-ocular reflex . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Central pathways ............................................
Ataxia and nystagmus in experimental cervical vertigo . . . . . . . . . . . . ..
Clinieal evidence for cervieal vertigo? . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Hypothetieal mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Differential diagnosis ........................................
Arthrokinetie nystagmus and self-motion sensation ..................
Other forms of nystagmus induced by non-vestibular stimulation ....
Postural imbalance with sensory polyneuropathy ...................
References ........................................................

441
442
442
443
443
443
444
444
445
446
446
447
447
449

Section K Psychogenic vertigo ...................................... 453


31

32

Psychiatrie dis orders and vertigo .................................... 455


Organic versus psychiatrie morbidity .............................
Vestibular dysfunction secondary to psychiatrie dis orders and
psychiatrie disorders secondary to vestibular dysfunction .........
How can psychogenic vertigo be diagnosed? .......................
Panic dis order ...................................................
Criteria for panic attack (DSM-IV 1994) ...........................
Agoraphobia ....................................................
Criteria for agoraphobia ........................................
Epidemiology ...............................................
Management ................................................
Acrophobia ......................................................
Psychotherapy for acrophobia and agoraphobia ....................
Psychogenic disorders of stance and gait ............................
Criteria for psychogenie disorders of stance and gait . . . . . . . . . . . . . . ..
Management ................................................
References ........................................................

456
457
458
459
459
459
459
460
460
461
461
462
463
466

Phobie postural vertigo

469

The clinieal syndrome

456

469

xxii

Contents

Aetiology and hypothetical mechanism .............................


Hypothetical mechanism: A disturbance of space constancy
due to decoupling of the efference-copy signal? . . . . . . . . . . . . . . . . . ..
Body sway in PPV ..............................................
PPV: A panic disorder? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Differential diagnosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Course and treatment .............................................
References ........................................................

470
471
472
473
475
475
478

Section L Physiological vertigo ...................................... 481


33

Motion sickness ................................................... 485

The clinical syndrome ............................................


Nausea and vomiting .............................................
Labyrinth function and motion sickness ............................
The sensory conflict theory (visual-vestibular mismatch) . . . . . . . . . . . . ..
Vestibular hyperexcitability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Incidence and susceptibility .......................................
Management: physical and medical prevention .......................
Physical prevention . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..
Visual prevention of motion sickness in vehicles ...................
Medical prevention ............................................
Space sickness ...................................................
References ........................................................

485
485
487
487
489
490
491
491
491
491
492
493

Subject Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 497

Glossary

Acrophobia: fe ar of heights; when "physiologieal height vertigo" induces a conditioned

phobie re action characterised by a dissociation of the objective and the subjective


risk of falling.
Adaptation: the adjustment of a sensory system to its environment or the process by
whieh this ability is achieved.
Agoraphobia: fear of wide open spaces or public places with excess anxiety, dizziness
and postural imbalance.
Alternobaric vertigo: transient vertigo due to pressure changes in the middle ear whieh
primarily affects divers and aircrew.
Antiemetics: drugs that control nausea and vomiting by acting on the medullary vomiting centre, the chemoreceptor trigger zone, or the gastrointestinal tract itself.
Antivertiginous drugs: vestibular suppressants, including anticholinergics, antihistamines, and benzodiazepines, which are used for symptomatic relief of distressing
vertigo, nausea and vomiting by downregulating vestibular excitability.
Arthrokinetic nystagmus: a purely somatosensory nystagmus and illusion of self-motion
with passive movements of the limbs in stationary subjects.
Barotrauma: signs and symptoms associated with exposure to alterations in ambient
pressure, either an increase (diving, pressure chamber, explosions) or a decrease (flying, altitude chamber), e.g. decompression sickness.
Basilar migraine: migrainous attacks with aura signs and symptoms within the vertebrobasilar territory.
Benign paroxysmal positioning vertigo (BPPV): most common form of vertigo caused by
canalolithiasis of the posterior semicircular canal, less often of the horizontal or anterior semicircular canal.
Benign paroxysmal vertigo in childhood (BPV): recurrent episodie vertigo in childhood as a
migraine equivalent.
.
Benign recurrent vertigo (BRV): recurrent episodie vertigo as a migraine equivalent in

adults.

Bilateral vestibular failure (BVF): bilateralloss of vestibular function with unsteadiness of


gait in the dark and oscillopsia associated with head movements.
Buoyancy hypothesis: positional nystagmus and/or vertigo with specific gravity differential between cupula and endolymph (e.g. positional alcohol nystagmus, positional
heavy water nystagmus).
Canalolithiasis: benign paroxysmal positioning vertigo and nystagmus caused by free-

floating heavy debris (otoconia) within the posterior or horizontal semicircular


canal.

xxiv

Glossary

Canal plugging: surgical plugging of single semicircular canals for treating canalolithiasis in rare cases of benign paroxysmal positioning vertigo.
Cawthorne-Cooksey exercises: vestibular exercises for rehabilitation and promotion of
vestibular compensation.
Cervical vertigo: to-and-fro vertigo and unsteadiness of gait induced by stimulation of,
or lesions in, neck afferents.
Cireularveetionllinearveetion: optokinetically induced perception of apparent selfmotion.
Cogan's syndrome: auto immune disease of young adults with interstitial keratitis and
audio-vestibular symptoms.
Coriolis effeet: spatial disorientation (with nausea) through cross-coupled accelerations, when the head is undergoing a rotation about one axis and is tilted about a second axis.
Cortical astasia: lateral postural imbalance (lateropulsion) and tilts of perceived vertical
with acute unilaterallesions of the parieto-insular vestibular cortex.
Cupulolithiasis: benign paroxysmal positional vertigo and nystagmus caused by heavy
debris settled on the cupula of the semicircular canal, transforming it from a transducer of angular acceleration into a transducer of linear acceleration.
Delayed endolymphatic hydrops: acquired types of endolymphatic hydrops which are
sometimes separated from "idiopathic" Meniere's disease.
Dix and Hallpike manoeuvre: positioning of patients with benign paroxysmal positioning
vertigo into a head-hanging position with the head turned.
Dizziness and light-headedness: typical presyncopal symptoms with orthostatic hypotension or cardiac arrhythmias, which also occur with hyperventilation syndrome, panic
attacks, metabolic hypoglycaemia, or drug intoxication.
Downbeat nystagmus: central disorder of the vertical vestibulo-ocular reflex in pitch
plane with downbeat nystagmus, oscillopsia and fore-aft postural imbalance.
Endolymphatic hydrops: enlargement and distortion of the endolymphatic compartment due to insufficient fluid resorption in the endolymphatic sac or from blockage
of the endolymphatic duct.
Epileptic nystagmus: ictal nystagmus induced by occipital or temporo-parieto-occipital
epileptogenic foci involving the vestibular, visual, or ocular motor cortices.
Falls in the elderly: significantly increased risk of falls with increasing age because of
multimorbidity and ageing.
Familial episodie ataxia type 1 (EA-l) or type 2 (EA-2): inherited channelopathies which
manifest as recurrent ataxia with or without vertigo.
Fixation suppression of the vestibulo-ocular reflex: suppression of vestibular nystagmus
during head acceleration by voluntary fixation of a stationary target.
Gait-ignition failure: inability to initiate and sustain locomotion with start-and-turn
hesitation, shuffling, and freezing, but relatively normal gait once locomotion is initiated.
Habituation: the simplest form of learning with gradual adaptation to a stimulus or the
environment, e.g. habituation to motion sickness stimuli on a ship within days.
Head-extension vertigo: physiological to-and-fro vertigo with head extension, particularIy with the eyes closed or when standing on an unstable support.

Glossary

xxv

Height vertigo: physiological "distance vertigo" through visual destabilisation of postural balance when the distance between the subject's eye and the visible stationary
surroundings becomes critically large.
Hyperviscosity syndrome and vertigo: pathological hyperviscosity of the blood associated
with polycythemia, hypergammaglobulinaemia, or Waldenstrm's macroglobulinaemia which may cause venous obstruction of the peripherallabyrinth.
Lateropulsion: irresistible lateral falls of patients with acute caudal brainstem lesions

(e.g. Wallenberg's syndrome), vestibular thalamic lesions (thalamic astasia), or


vestibular cortex lesions ("pusher syndrome").
Liberatory manouevres: rapid positionings of patients with benign paroxysmal positioning vertigo (Brandt-Daroff exercises, Semont's manoeuvre, Epley's manoeuvre)
designed to free the semicircular canal of the heavy dot formed during canalolithiasis.
Mal de debarquement syndrome: persisting sensations of swinging, swaying, unsteadi-

ness and disequilibrium experienced after sea travel immediately upon disembarking.
Meniere's disease: endolymphatic hydrops with the dassic triad of fluctuating hearing

loss, tinnitus and episodic vertigo.


Mismatch concept: motion sickness or vertigo generated by an acute sensorimotor conflict (mismatch between the converging sensory inputs or between the expected and
actually perceived sensory pattern).
Mondini dysplasia: malformation of the membranous and osseous labyrinth with combined auditory and vestibulary loss.
Motion sickness: distressing syndrome with nausea and vomiting induced by unfamil-

iar body accelerations in vehides to which the person has not adapted or by intersensory mismatch involving conflicting visual and vestibular stimuli.
Neural integrator or gaze-holding function: neural process that integrates velo city to position in order to hold gaze steady at the end of an eye movement in an eccentric position of the orbit when elastic forces tend to return it to primary position.
Ocular tilt reaction: disorder of the vestibulo-ocular reflex in roll; eye-head synkinesis

consisting of lateral head tiIt, skew deviation, and ocular torsion; VOR-OTR with pontomedullary lesions of the vestibular nudei, which subserve the vestibulo-ocular
reflex; integrator-OTR with lesions of the rostral midbrain integration centre for eyehead co ordination in roll plane.
Optokinetic motion sickness: symptoms of motion sickness when viewing large moving
visual scenes (simulator sickness).
Oscillopsia: apparent movement of the visual scene due to involuntary retinal slip in
patients with acquired ocular oscillations or deficient vestibulo-ocular reflex.
Otolithic vertigo: otolith dysfunction causing non-rotatory, to-and-fro vertigo associated
with head acceleration and unsteadiness of gait, oscillopsia, perceived tiIt and ocular
deviation.
Ototoxic agents: drugs and substances that transiently or permanently damage the

cochlea or the vestibular labyrinth.


Paroxysmal dysarthria/ataxia: non-epileptic manifestation of paroxysmal attacks in mul-

tiple sderosis by ephaptic activation of adjacent demyelinated axons.

xxvi

Glossary

Perilymph fistula: rupture of the otie capsule, usually at the oval or the round window,
whieh causes perilymph leakage and abnormal transfer of pressure changes.
Phobie postural vertigo: frequent psychosomatie postural vertigo with unsteadiness of
gait distinguishable from agoraphobia and acrophobia.
Pitch: sagittal plane of action of the vestibulo-ocular reflex with nead extension or
flexion about the binaural horizontal y-axis.
Plastieity of the vestibular system: mechanisms including habituation and readjustment
to new environmental conditions or central compensation and sensory substitution
for vestibular deficits.
Positional alcohol nystagmus (PAN): direction-changing, positional nystagmus and vertigo as a result of alcohol intoxieation, secondary to specific gravity differential
between cupula and endolymph (buoyancy hypothesis).
Positional vertigo: vertigo induced by changes in head position relative to the gravitational vector; in positioning vertigo head movement rather than head position is the
precipitating factor.
Positioning vomiting: vestibulo-autonomie central positioning vomiting due to lesions
between the vestibular nuclei and the archicerebellar vermis.
Pressure fistula test: also known as the Hennebert sign; when pressure changes within
the external auditory canal evoke ocular deviation, nystagmus, oscillopsia, vertigo, or
postural imbalance in patients with perilymph fistula.
Purkinje effect: tumbling sensation of turning about an off-vertical body axis when the
head is tilted during a post-rotational semicircular canal response.
Ramsay Hunt syndrome: herpes zoster otieus.
Roll: frontal plane of action of the vestibulo-ocular reflex with head motion in roll
about the line of sight (x-axis).
Room-tilt illusion: transient upside-down vision or 90 tilt due to an acute vestibular
tone imbalance whieh elicits a transient cortieal mismatch between the visual and
vestibular 3D-coordinate maps.
Rotational vertebral artery occlusion: transient ischemic attacks with vertigo, nystagmus,
and ataxia secondary to vertebral artery compression with rotation al head motion.
Rotatory vertigo: vertigo occurring with acute unilateral peripheralloss of vestibular
function, pontomedullary brainstem lesions near the vestibular nuclei, or paroxysmal
stimulation of these structures.
Scheibe syndrome: cochleo-saccular malformation with sparing of the semicircular
canals and the utricle.
Semicircular canal vertigo: typieal signs and symptoms of which are rotational vertigo
and deviation of perceived straight-ahead, spontaneous vestibular nystagmus with
oscillopsia, postural imbalance with Romberg fall, and nausea and vomiting if severe.
Senile gait: cautious gait of older people.
Simulator siekness: motion sickness elicited in high-fidelity visual simulators or virtual
environment systems.
Skew deviation (skew-torsion sign): vertical misalignment of the visual axes due to a gravieeptive vestibular tone imbalance in roll plane.
Space constancy mechanism: adequate perception of a stable world despite visual motion
stimulation, eye-head motion, or locomotion.

Glossary

xxvii

Space sickness: motion sickness in microgravitational environments elicited by head

movements.
Spatial hemineglect: impairment of focal visuo-spatial attention toward the contralater-

al side of lesions of the inferior parietallobule or frontal premotor cortex, also involving the vestibular system.
Thalamic astasia: lateral postural imbalance (lateropulsion) in acute vestibular thalamic
lesions without motor weakness, sensory loss, or cerebellar signs.
Traumatic otolithic vertigo: traumatic dislocation of otoconias resulting in unequalloads
on the macular beds and causing transient head motion intolerance (oscillopsia and
postural imbalance).
Tullio phenomenon: pathological sound-induced vestibular signs and symptoms in
patients with perilymph fistula.
Tumarkin's otolithic crisis: vestibular drop attacks in Meniere's disease.
Upbeat nystagmus: central dis order of the vertical vestibulo-ocular reflex in pitch with
upbeat nystagmus, oscillopsia, and postural imbalance.
Vascular fistula test: test for bilateral compression of the jugular vein which causes eye
movements or vertigo in patients with perilymph fistula.
Vestibular atelectasis: collapse of the walls of the ampulla and utricle with unilateral or
bilateral vestibular dysfunction.
Vestibular compensation: central readjustment of a lesion-induced vestibular tone
imbalance; it consists of multiple processes for perceptual, vestibulo-ocular, and
vestibulospinal readjustments, which have different time courses and occur at different sites in the brain and spinal cord.
Vestibular cortex: multiple multisensory temporoparietal cortex areas including the
parieto-insular vestibular cortex, areas 2v, 3aV, 6, and 7.
Vestibular drop attacks: sudden falls due to vestibulospinalloss of postural tone caused
by inadequate otolithic stimulation in Meniere's disease.
Vestibular epilepsy: episodic vertigo secondary to focal discharges from the vestibular
cortex.
Vestibular exercises: physical therapy for vestibular rehabilitation to readjust vestibuloocular and vestibulospinal reflexes or to promote central habituation so as to prevent
motion sickness.
Vestibular falls: peripheral or central vestibular dysfunction causing irresistible or
unexpected falls.
Vestibular neurectomy: transeetion of the vestibular nerve in rare cases of intractable
labyrinthine vertigo, particularly in severe cases of Meniere's disease.
Vestibular neuritis (VN): acute partial unilateral vestibular loss due to inflammation of
the vestibular nerve with rotatory vertigo, nystagmus, postural imbalance, nausea and
vomiting.
Vestibular paroxysmia (disabling positional vertigo): paroxysmal vertigo, oscillopsia, tin-

nitus and postural imbalance due to neurovascular cross-compression of the VIIIth


nerve.
Vestibular substitution: process by which parts of the missing vestibular inputs for, e.g.

perceptual, ocular motor, and postural control are substituted by vision or cervical
proprioception.

xxviii

Glossary

Vestibulo-autonomie regulation: functions involving respiratory and cardiovascular contral with changes in body position, affective and emotional responses with body
accelerations, nausea and vomiting, and modulation of sleep.
Vestibulocollie reflex (VCR): a three-neuron reflex arc from vestibular afferents to neck
motor neurons to stabilise the head position in space.
Vestibulogenic epilepsy: variety of seizures induced by vestibular stimulation or dysfunction.
Vestibulo-ocular reflex (VOR): a three-neuron reflex arc that serves to hold constant the

direction of gaze in space during head movements by moving the eyes in the direction opposite to that of the head with a velocity and amplitude wh ich compensate for
the head motion.
Vestibulospinal reflexes: phasic and tonic reflexes that stabilise head and upright posture

in relation to gravity via mediation of lateral vestibulospinal, medial vestibulospinal,


and reticulospinal tracts.
Visual cliff phenomenon: innate visual depth avoidance without former experience of
falling off edges.
Visual vertigo: spatial disorientation, misperception of motion, and postural imbalance
induced by unusual visual stimulation or visual dysfunction.
Volvular epilepsy: rare ratatory seizures characterised by paroxysmal repetitive walking

in small circles.
Wallenberg's syndrome: dorsolateral medullary infarction with involvement of the
vestibular nuclei causing lateropulsion of the eyes and the body.
Yaw: horizontal plane of action of the vestibulo-ocular reflex with head rotations
about the vertical z-axis.

SECTION A
Vertigo: symptoms, syndromes,
disorders

Introduction

three sensory systems subserve both static and


dynamic spatial orientation, loeomotion, and control
of posture by constantly providing reafferent cues.
The sensory information is partially redundant in
that two or three senses may simultaneously provide
similar information about the same action. Thanks
to this overlapping of their funetional ranges, it is
possible for one sense to substitute, at least in part,
for deficiencies in the others. When information
from two sensory sources conflicts, the intensity of
the vertigo is a function of the degree of mismatch;
it is increased if information from an intact sensory
system is lost, as for example in a patient with pathological vestibular vertigo who closes his eyes. The
distressing sensorimotor consequences of the mismatch are frequently based on our earlier experiences with orientation, balance, and locomotion, i.e.
there is amismatch between the expected and the
actually perceived pattern of multisensory input.
Vertigo may thus be induced by physiological
stimulation of the intact sensorimotor systems
(height vertigo; motion sickness) or by pathological
dysfunction of any of the stabilising sensory
systems, especially the vestibular system (Table 1.1).
The symptoms of vertigo include sensory qualities
identified as arising from vestibular, visual, and
somatosensory sources. As distinct from one's perception of self-motion during naturallocomotion,
the experience of vertigo is linked to impaired perception of a stationary environment; this perception
is mediated by central nervous system processes
known as "space constancy mechanisms". Loss of
the external stationary reference system required for
orientation and postural regulation contributes to
the distressing mixture of self-motion and surround
motion (Brandt and Daroff 1980).

Vertigo is an unpleasant distortion of statie gravitational orientation, or an erroneous pereeption of


motion of either the sufferer or the environment. It
is not a disease entity, but rather the outeome of
many pathological or physiologieal processes. Vertigo
is best deseribed as a multisensory and sensorimotor
syndrome with pereeptual, postural, oeular motor,
and autonomie manifestations indueed by either

unusual and therefore unadapted (motion) stimulation of the intaet sensory systems, or
pathological (lesional) dysfunetion.

Vertigo, dizziness, and disequilibrium are common


eomplaints of patients of all ages, particularly the
elderly. As presenting symptoms, they oeeur in
5-10% of all patients seen by general praetitioners
and 10-20% of all patients seen by neurologists and
otolaryngologists. The elinieal speetrum of vertigo is
broad, extending from vestibular rota tory vertigo
with nausea and vomiting to presyneope lightheadedness, from drug intoxieation to hypoglycaemie
dizziness, from visual vertigo to phobias and panie
attaeks, and from motion sickness to height vertigo.
Appropriate preventions and treatments differ for
different types of dizziness and vertigo; they include
drug therapy, physical therapy, psyehotherapy, and
surgery.
The following introduetion to this monograph
will foeus on vestibular syndromes as the basic
model for understanding the structural and functional aspects of disorders whose major complaint is
vertigo, dizziness, or disequilibrium. The vestibular
system is part of a multisensory and sensorimotor
network for spatial orientation and balance eontro!.

The "vestibular" vertigo syndromes

Table 1.1. Physiological or pathological vertigo

Vertigo usually implies amismatch between the


vestibular, visual, and somatosensory systems. These

Pathological dysfunction

Physiological stimulation

Height vertigo
Motion sickness
Labyrinthine and vestibular nerve disorders
Central vestibular disorders

Vertigo

Signs and symptoms

pathways (p. 18) to aetivate the medullary vomiting eentre (p. 486).

Physiologieal and clinieal vertigo syndromes (Table


1.2) are eommonly eharaeterised by a eombination
of phenomena involving pereeptual, oeular motor,
postural, and autonomie manifestations: vertigo,
nystagmus, ataxia, and nausea (Fig. 1.1; Brandt and
Daroff 1980). These four manifestations eorrelate
with different aspeets of vestibular funetion and
emanate from different sites within the eentral
nervous system.
e
e

The vertigo itself results from a disturbanee of


eortieal spatial orientation.
Nystagmus is seeondary to a direetion-speeifie
imbalanee in the vestibulo-oeular reflex, whieh
aetivates brainstem neuronal eireuitry.
Vestibular ataxia and postural imbalanee are
eaused by inappropriate or abnormal aetivation
of monosynaptie and polysynaptie vestibulospinal pathways.
The unpleasant autonomie responses with
nausea, vomiting, and anxiety travel along
aseending and deseending vestibulo-autonomie

Perlpheral
labyrlnthlne
le.lon

Optokinetic

Somatoklnetic

Syndromal manifestations of vertigo

Syndrome

Manifestation

Pereeptual
Ocular motor
Postural
Autonomie

Vertigo, disorientation
Nystagmus, ocular deviation
Ataxia, falls
Nausea, vomiting, anxiety

Under eertain conditions, distressing symptoms


and malaise may be preeeded by a pleasurable autonomie sensation, whieh is presumably mediated
through the limbie system and aeeounts for the popularity of amusement park rides and the like.

The mismatch concept


Physiologieal vertigo (motion siekness; p. 485) and
pathologieal vertigo (peripheral or eentral vestibular
dysfunetion) are thought to be generated by an aeute

VESTIBULAR
FUNCTION

VERTIGO
SYNDROME

Spatial orientatlon
Motion perceptio n

VERTIGO

Vestibulo - Ocular
Ref lex

NYSTAGMUS

Posture

ATAXIA

PATHOLOGICAL
VERTIGO

PHYSIOLOGICAL
VERTIGO

Vestibular

lable 1.2.

I
I
"I,$' I- (
I~- I

[][]

Vestlbular
epllepsy

Perlpheral
elghth nerve
le.lon
Cenlral
vestlbular
lulon

/'

Paneto temporal
Cortex

/'

I
~ "-

cantral
veslibular
palhways

Brainstem

Spinal

Madullary
vomiting
centre
Limbic
system

Vegetative
effects

NAUSEA

Fig.1.1. Classification of physiological vertigo and vestibular disorders with their origin at different sites within peripheral or central
vestibular structures. Vestibular disorders are not clinical entities but different sensorimotor syndromes arising from unusual stimulation or lesional dysfunction. (From Brandt and Daroff 1980.)

Introduction

sensorimotor conflict (mismatch) between the converging sensory inputs and the expected sensory
patterns (Fig. 1.2) or a vestibular tone imbalance (p.
73). Amismatch arises, for example, when the multisensory consequences of being a passenger in a
moving vehicle or of moving actively do not match
the expected patterns which have been calibrated by
prior experience of active locomotion (p. 487). Thus,
it is the sensory mismatch (e.g. visual-vestibular or
between right and left vestibular input) rather than
the sensory loss which causes vertigo. The absence of
one channel of the redundant sensory input important as it is for dem an ding balancing tasks in
sports - rarely manifests as vertigo. Inappropriate
information from one or multiple sensory systems
produces an illusion of body motion and causes
vertigo. An acute unilaterallabyrinthine dysfunction
(see vestibular neuritis; p. 67) causes vertigo because

the sensation of self-motion induced by the vestibular tone imbalance is contradicted by vision and the
somatosensors.

The vestibulo-ocular reflex (VOR)


It is possible to recognise faces and to read while
walking because the vestibulo-ocular reflex (VOR)
compensates (when it is functioning correctly) for
the high-frequency head perturbations. The VOR
normally serves to hold constant the direction of
gaze in space during head movements. It achieves
this by moving the eyes in the direction opposite to
that of the head, with a velo city and amplitude which
"compensates" for the head motion. If the amplitude

--->-----------~----------

corollary
discharge

IRE-AFFERENCES

'"
voluntary
motion

'"-

EXPECTED
AFFERENCES

DJ
DJ

CENTRAL
STORE

rn

---oe:

comparison

I
DJ
DJ

\
DJ
1 2]

rn rn
space constancy

habituation

--::---

mismatch

.J.

vertigo

Fig.1.2. Schematic diagram of the sensory conflict or the neural mismatch concept of vertigo and motion sickness. An active movement leads to stimulation of the sensory organs whose messages are compared with a multisensory pattern of expectation calibrated
by earlier experience of motions (central store). The pattern of expectation is prepared either by the efference copy signal which is
emitted parallel to and simultaneously with the motion impulse, or by vestibular excitation during passive transportation in vehicles.lf
concurrent sensory stimulation and the pattern of expectation are in agreement, self-motion is perceived while "space constancy" is
maintained.lf, for example, there is no appropriate visual report of motion, as a result of the field of view being filled with stationary
environmental contrasts (reading in the car). a sensory mismatch occurs. With repeated stimulation, motion sickness is induced
through summation; the repeated stimulation leads to arearrangement of the stored pattern of expectation, however, so that a habituation to the initially challenging stimulation is attained within a few days. An acute unilaterallabyrinthine 1055 causes vertigo,
because the self-motion sensation induced by the vestibular tone imbalance is contradicted by vision and the somatosensors.

Vertigo

and/or velocity of eye movements are inappropriate,


the resultant shift in the direction of gaze causes a
displacement or slip of the retinal image which may
be perceived as an apparent motion of the fixated
visual scene, i.e. oscillopsia (Wist et al. 1983).
Compensatory eye movements, which are initiated
by head movements, make use of input from

the semicircular canals and the otoliths (VOR),


the retina (optokinetic reflex, OKN), and
neck somatosensors (cervico-ocular reflex, COR).

These different loops provide

converging and redundant information ("functional overlapping"),


partial compensation for each other's deficiencies
("functional substitution"), and
preferred frequency ranges of action ("functional
specialisation") .

High frequency perturbations exceeding 3 Hz (with


peak head velocities of up to 150 deg/s), which occur
during naturallocomotion (Snyder and King 1988;
King et al. 1992), are the domain of the rapid threeneuron VOR with its short latency of action of about
16 ms. Similar latencies have also been demonstrated
for otolith reflexes affecting eye position (Dieterich
et al. 1989). When the VOR generates eye rotations
that compensate for head movements, the gain of the
response (eye movement/head movement) is 1.0.
Most laboratory studies on the human VOR, however, report that the gain of the reflex is 0.75 or less
(Collewijn 1989; Leigh and Brandt 1993). This has
led to the conviction that the VOR, on its own, cannot provide clear and stable vision during head
movements but requires inputs from the visual system and neck proprioception (for review see Leigh
and Zee 1998).
The VOR has three major planes of action (Sect.
C, Fig. 1; p. 169):

horizontal head rotation about the vertical zaxis,yaw


head extension and flexion about the horizontal
y-axis, pitch
lateral head tilt ab out the horizontal x-axis, roll

These planes represent three-dimensional space in


the vestibular and oculomotor systems (Cohen and
Henn 1988) and make different demands upon the
VOR to ensure perception of a stable world ("space
constancy mechanism") during head rotations in
yaw, pitch or roll. The VOR in roll is mediated by the
vertical anterior and posterior semicircular canals;
but with sustained head tilt, static otolith reflexes
partially compensate for eye position in roll (ocular

counter-rolling). From a visual standpoint, the torsional VOR need not be as efficient as its horizontal
or vertical counterparts, since head movements in
roll do not displace images from the fovea (Leigh
and Zee 1998). Only in the periphery of the retina
(areas of sparcer photoreceptor density) will there
be an appreciable slip of images in the absence of
compensatory eye movements. Certain torsional disparities are weIl tolerated by visual processing
mechanisms (Bishop 1978; Kertesz 1983; Dieterich
and Brandt 1992) (e.g. patients with Wallenberg's
syndrome seldom complain of torsional diplopia),
and the stability of torsional gaze, although much
less constant than horizontal or vertical gaze (Ott et
al. 1992), does not appear to impair visual acuity or
perception. The VOR has different properties in the
torsional plane as opposed to those of the horizontal
and vertical planes (Ferman et al. 1987; Leigh et al.
1989; Seidman and Leigh 1989). The gain of the torsional VOR under optimal circumstances is never
high enough to compensate for natural head movements (typically 0.65). Moreover, gaze stability (Ott
et al. 1992) and the dynamic properties ofVOR during head rotation in roll differ, and the torsional
optokinetic response is weak (Collewijn et al. 1985).
During locomotion translations of the head occur
due to head perturbations and forward motion
through the environment (Schwarz et al. 1989; Paige
1989; Hess and Dieringer 1991). The component of
the VOR wh ich responds to head translations
depends on the otolithic organs, which are switched
on when the subject views a near object (Viirre et al.
1986; Schwarz et al. 1989; Paige 1989). This conceptualisation has led to the development of tests for
otolithic function (Gresty and Bronstein 1992;
Gresty et al. 1992), for example, comparison of the
magnitude of eye movements during fixation of near
and distant targets when the subject is translated laterally or fore-and-aft on a parallel swing (a swing
with rigid vertical bars prevents angular motion)
(Baloh et al. 1988). Another technique consists in
placing the head of the subject in front of the axis of
rotation of a vestibular chair; this achieves a combined angular-linear movement that stimulates both
the semicircular canals and otoliths. The effect of
otolith stimulation can also be measured during sustained rotation about an axis tilted from earth vertical (off-vertical axis rotation, OVAR) such as
"barbecue spit" rotation (Guedry 1965; Darlot et al.
1988; Wall and Furman 1989).
The spatial planes of the semicircular canals
roughly represent the planes of the lines of action of
the extraocular muscles (Fig. 1.3). Spatial organisation of the right and left labyrinths in the temporal
bones is such that horizontal and vertical semicircular canals can be paired with respect to their optimal

Introduction
plane of function, although three-dimensional
reconstruction of semicircular canals did not show
this to be perfect in the individual (Takagi et al.
1989). Hence, for example, ampullofugal stimulation
(excitation) of the anterior semicircular canal is
associated with ampullopetal stimulation (inhibition) of the opposite posterior semicircular canal.
The purpose of the optokinetic and pursuit systems
is to maintain fovealisation of visual targets during
low frequency movements of the target or head. The
visual pursuit system is able to suppress the vestibulo-ocular reflex during head acceleration (fixation
suppression of the VOR) and - by virtue of its optokinetic after-nystagmus ("storage mechanism") - to
counterbalance undesired post-rotational VOR when
the head movements cease (deceleration). Thus, the
visual system helps to suppress post-rotational nystagmus by counterbalancing vestibular activity with
optokinetic activity in the opposite direction (Barret
and Hood 1988).
A lesional tone imbalance between the two corresponding semicircular canals of the right and left
labyrinth results in linear or rotary ocular deviation
or spontaneous nystagmus, with a directional preponderance of the VOR gain independent of the type
of vestibular stimulation (head rotation or thermal
irrigation). Spontaneous nystagmus can be suppressed by fixation. It should be observed, therefore,
when Frenzel's glasses are used or during ophthalmoscopy (Zee 1978). In chronic unilaterallesions,
when spontaneous nystagmus has disappeared due
to central compensation, head-shaking nystagmus
still demonstrates clinically the asymmetry of the
velocity storage (Takahashi et al. 1990). In this case a
transient nystagmus can be seen with Frenzel's
glasses, which is elicited by 10 s of vigorous head
shaking. This method cannot distinguish clearly
between peripheral and centrallesions, but differentiation is possible with the VOR-bedside test (p. 39),
as described by Halmagyi and Curthoys (1988).

Neuronal network of the VOR


The neuronal network mediating VOR from the horizontal and vertical canals is based on sensory convergence within a three-neuron reflex arc (Fig. 1.3).
It links a set of extraocular muscles with their primary action aligned to the particular spatial plane of
the horizontal, the anterior, or the posterior canal.
Sensorimotor transformation from canal planes to
the planes of eye movements has been demonstrated
at the level of second-order neurons within the
vestibular nuclei. These neurons, projecting to ocular motor neuron pools, always contact their two
respective eye actuators in the excitatory as well as

I
1
1

1
1

1--t -. +

: I
L.

LR

I
I
I

111
.................... -I
... 1".........................
..; VI

23

41

\-1

IV

'

\1
"

~I
"

HC
;I'
;I'

PC

~I
56

1
1
1
I

AC
HC
PC

Fig. 1.3. Schematic representation of the three-neuron


vestibulo-ocular reflex from semicircular canals to extraocular
eye muscles (excitatory pathways only). Sensorimotor transformation occurs from canal planes to the planes of eye movements
in such a way that the neurons always contact their two respective extraocular eye muscles. Horizontal canal pathways contact
only abducens motor neurons to activate the ipsilateral medial
rectus and the contralaterallateral rectus muscle. Anterior canal
pathways contact ipsilateral superior rectus and contra lateral
inferior oblique muscles. Posterior canal pathways contact ipsilateral superior oblique and contra lateral inferior rectus muscles.
Perceptual and vestibulospinal pathways are omitted. AC: anterior
canal ; HC: horizontal canal; PC: posterior canal; SR: superior
rectus muscle; MR: medial rectus muscle; SO: superior oblique
muscle; 10: inferior oblique muscle; IR: inferior rectus muscle; LR:
lateral rectus muscle; 111: oculomotor nucleus; IV: trochlear nucleus;
VI: abducens nucleus. (Modified after Graf and Simpson 1981 and
Graf and Brunken 1984.)

Vertigo

in the inhibitory vestibulo-ocular motor link (pushpull operational mode) (Graf et al. 1983; Graf and
Ezure 1986). The excitatory pathways of the VOR
cross the midline of the pontomedullary brainstem
tegmentum; the inhibitory pathways ascend ipsilaterally. Excitation of the horizontal semicircular
canals causes horizontal deviation of the eyes in
yaw; excitation of the anterior semicircular canals
causes upward deviation of the eyes (see downbeat
nystagmus; p. 199); excitation of the posterior semicircular canals causes downward deviation of the
eyes (see upbeat nystagmus; p. 205).
Although oligosynaptic pathways are essential for

the short-latency properties of the VOR, they represent only a portion of the connections subserving
this reflex (Leigh and Zee 1998; Leigh and Brandt
1993). Other pathways are needed to generate appropriately calibrated eye movements; to account, for
example, for the proximity of the object of regard
(otolithic and visual inputs); and to hold gaze steady
at the end of the compensatory eye movement (gazeholding or neural integrator function). Furthermore,
primary vestibular inputs that serve the VOR send
axon collaterals to neurons responsible for vestibulospinal reflexes and to cortical areas involved in the
perception of self-motion (Fig. 1.4).

PERCEPTION

thalamus I
cortex

ri MLF

INe

1~-,--,-Eii7 ~
so

VESTIBULOOCULAR
REFLEX
mesencephalic

IV

MLF

pontomedullary

VESTIBULO
SPINAL
REFLEX

spinal cord

Fig.1.4. Schematic representations of the vestibulo-ocular reflex. Elements that contribute to the overall, sensorimotor vestibular
response. Inputs from the horizontal (HC), anterior (AC), and posterior (PC) semicircular canals converge with otolithic, visual, and
somatosensory afferents in the vestibular nuclear complex (VN). The outputs from the neural network in the VN contact the extraocular muscles; here the principal three-neuron arc connections of the PC are shown passing to the trochlear (IV) and oculomotor nuclei
(111) wh ich contact the superior oblique and inferior rectus muscles.ln addition, connections from the AC and PC contact the interstitial
nucleus of Cajal (lNCl, wh ich is important for eye-head coordination in roll and in vertical gaze-holding, and to the rostral interstitial
nucleus of the medial longitudinal fasciculus (riMLFl, which is important in generating quick phases of vestibular nystagmus in the
vertical and torsional planes. (Divergence or convergence of the vestibular nuclei network is not shown.) The VN output also projects
to the spinal cord, to generate vestibulospinal reflexes, and to thalamus and cortex, to provide inputs for perception of movement.
Thus, VOR pathways also mediate posture and perception. (From Leigh and Brandt 1993.)

Introd uction

The vestibular nuclei can be conceptualised as a


neural network capable of"learning" (Anastasio and
Robinson 1989). For example, if the VOR is to compensate for head rotation in the vertical (pitch)
plane, signals from the anterior and posterior semicircular canals must be transformed to produce
appropriate contractions of the extraocular muscles
- the vertical rectus and oblique muscles; this transformation is necessary since the co ordinate frames
of the canals and the muscles differ (Anastasio and
Robinson 1990). Experimental data indicate that
directions of head rotation for which neurons in the
vestibular nucleus show their maximum response
(sensitivity vector) vary widely and do not show any
uniformity of alignment (Baker et al. 1984). This
finding suggests that signals from different semicircular canals converge on these neurons. When a
three-layered neural network was successfully
"trained" to solve the canal-muscle transformation
problem, the sensitivity vectors of the middle layer
of the network were widely dispersed, similarly to
those recorded experimentally (Anastasio and
Robinson 1990). This approach - distributed parallel
processing of neural signals - has gained considerable popularity and led to a number of insights into
the way that the brain makes its computations. For
example, there is no unique solution to problems like
the canal-muscle transformation problem, but the
relative weighting between different synapses of the
model is determined as the "modellearns" to "solve
the problem" (Anastasio and Robinson 1990).
Further, the dynamic properties of the VOR of the
newborn differ from those of adults - this maturation occurs during the first two months of life as
visual information becomes available to provide the
"error signal" necessary to train a network of
neurons (Weissman et al. 1989).
Another important function performed by
neurons in the medial vestibular nucleus and the
adjacent nucleus praepositus hypoglossi is the gazeholding or "neural integrator" function. Most ocular
motor commands are encoded in terms of velo city,
but a position component is also necessary to hold
the eye in an excentric position in the orbit against
elastic forces that tend to return it to primary position. Destruction of the medial vestibular nucleus
and nucleus praepositus hypoglossi abolishes all
gaze-holding function in the horizontal plane
(Cannon and Robinson 1987). In the vertical plane,
the interstitial nucleus of Cajal contributes to gazeholding (Crawford et al. 1991). A neural network
approach has also been used to determine how the
vestibular nucleus achieves gaze-holding (neural
integrator) properties and has provided some interesting insights into the effects of lesions on this network (Cannon and Robinson 1985; Arnold and

Robinson 1991) and the manner of its recovery from


a peripheral vestibular lesion (Anastasio 1992).

VOR mediation of perception and postural


adjustments
Variations in the structural elements and functional
characteristics of the VOR under certain stimulus
conditions have been thoroughly investigated by
ocular motor scientists. The VOR gain (eye versus
head velocity), the phase relationship between stimulation and eye movement (input-output), the time
constant of the VOR due to a combination of
mechanical properties of the cupula-endolymph
system (acting as a torsion pendulum), and the central storage mechanism are well known. However,
the commonly held view that the VOR merely serves
to stabilise gaze in space is too simple a concept. Its
neuronal pathways also include ascending input to
thalamocortical projections for perception (Fig. 1.4)
as well as descending input to vestibulospinal projections for adjustments of head (vestibulocollic
reflex, p. 10) and body posture (vestibulospinal
reflexes; p. 10). This means that both physiological
stimulation and pathological dysfunction of the
semicircular canals not only provoke nystagmus but
inevitably cause a direction-specific concurrent rotatory vertigo and postural imbalance. Disorders of
posterior semicircular canal function, for example,
clearly demonstrate the close functional link
between perception, eye movements, and posture:
e.g. canalolithiasis of the posterior semicircular
canal (benign paroxysmal positioning vertigo;
p. 251) causes paroxysms of nystagmus, rotatory vertigo, and postural imbalance precipitated by rapid
changes in head position. Another example is disinhibition of bilateral posterior canal tone due to
dysfunction of the flocculus which causes the downbeat nystagmus/vertigo syndrome with a tendency
to fall backwards (p. 201).
Finally, the VOR is also involved in object localisation in space, particularly in the absence of an external (visual, auditory, haptic) reference. Maurer et al.
(1997) developed a concept, "according to which
humans evaluate the kinematic state of a visual
object in space by (a) relating it to that of the body
support by means of an essentially ideal proprioceptive co ordinate transformation and (b) relating, in
turn, the kinematic state of the support to a vestibularly derived notion of space, using a proprioceptive
coordinate transformation that 'knows' the vestibular transfer characteristics."
For details on the plasticity of the VOR (p. 55),
vestibular compensation and substitution (p. 56),
and transmitters (p. 58), see Chap. 3. The purpose,

Vertigo

10

properties, neural substrate, and dis orders of the


VOR can be summarised as follows (Leigh and
Brandt 1993):
"Conventional views of the vestibulo-ocular reflex
(VOR) have emphasised testing with caloric stimuli and by passively rotating patients at low
frequencies in achair. The properties of the VOR
tested under these conditions differ from the performance of this reflex during the natural function for which it evolved - locomotion. Only the
VOR (and not visually mediated eye movements)
can cope with the high-frequency angular and
linear perturbations of the head that occur during
locomotion; this is achieved by generating
eye movements at short latency 16 msec).
Interpretation of vestibular testing is enhanced by
the realisation that, although the di- and trisynaptic components of the VOR are essential for this
short-latency response, the overall accuracy and
plasticity of the VOR depend upon a distributed,
parallel network of neurons involving the vestibular nuclei. Neurons in this network variously
encode inputs from the labyrinthine semicircular
canals and otoliths, as well as from the visual and
somatosensory systems. The central vestibular
pathways branch to contact vestibular cortex (for
perception) and the spinal cord (for control of
posture). Thus, the vestibular nuclei basically
co ordinate the stabilisation of gaze and posture,
and contribute to the perception of verticality and
self-motion. Consequently, brainstem dis orders
that disrupt the VOR cause not just only nystagmus, but also instability of posture (e.g. increased
fore-aft sway in patients with downbeat nystagmus) and disturbance of spatial orientation (e.g.
tilt of the subjective visual vertical in Wallenberg's
syndrome)."

Vestibulocollic reflex
Stabilisation of the head in space is required not
only for adequate motor performance, such as maintaining balance while standing or walking, but also
for adequate perception of head-fixed sensory information, such as visual and auditory inputs (Wilson
et al. 1995). The basic circuitry of the vestibulocollic
reflex, which collaborates with the cervicocollic
reflex, is a three-neuron arc from vestibular afferents
with a vestibulocollic interneuron to the neck motor
neurons. It has been hypothesised that a steady angle
of head orientation in the sagittal plane with respect
to gravity may be necessary to optimise the sensitivity of the otolithic organs in order to sense linear
accelerations (Pozza et al. 1990, 1991). The vestibular

system seems to work in a "top-down" fashion, the


brain paying particular attention to the orientation
of the head in space rather than to the orientation of
the trunk ("bottom-up" from feet to head), to which
the head is "rigidly" attached. Normal subjects
express a preference for the former organisation,
since it allows the head to serve as a "gaze-anchored"
reference system. However, in patients who have lost
vestibular sense and in children learning to walk, a
"bottom-up" control of posture may prevail, with a
tendency to anchor the head to the trunk during
locomotion (Pozza et al. 1991; Assaiante and
Amblard 1990). Measurements in patients with loss
of peripheral vestibular function showed that head
rotations during walking or running in place are not
increased in angular velo city or frequency. These
patients can stabilise their head during free walking
as well as normal subjects (Grossman and Leigh
1990; Leigh et al. 1992; Pozza et al. 1990, 1991). Thus,
mechanical properties, due to neck muscle tone and
the inertia of the cranium, prevent excessive vibrations of the head after the loss of vestibular sense;
however, these patients have lost the ability to sustain a steady angle of head orientation in the sagittal
plane (standard deviation of 3 deg in normal subjects, Pozza et al. 1991).

Vestibulospinal reflexes
Phasic and tonic vestibulospinal reflexes act to stabilise head and upright posture in relation to gravity.
Stimulation of canal or otolith receptors leads to a
variety of patterns of activation of neck and body
muscles, all tending to prevent falling and to maintain the position assumed (Wilson and Peterson
1978). Short-latency phasic reflexes are mediated by
the semicircular canals, largely via the medial
vestibulospinal (collic) tract that links each of the
semicircular canals to a set of neck muscles to
stabilise head position (vestibulocollic reflex) in
space (Markharn 1987; Shinoda et al. 1988). The
combination of vertical and ipsilateral horizontal
semicircular canal input on many secondary medial
vestibulospinal tract neurons suggests a contribution to the vestibulocollic reflex (Iwamoto et al.
1996). Stimulation of a semicircular canal evokes
head and body movements, which parallel the plane
of the particular canal. The patterns of semicircular
canal input to neck motor neurons are closely related
to the mechanical actions of the individual neck
muscles and the optimal stimulus to the semicircular canal. As a result the connections will te nd to
stabilise head positions in response to head pertur-

11

Introduction

bations (Shinoda et al. 1997). Individual neck motor


neurons are connected to both labyrinths, but not to
the same semicircular canal in each (Wilson and
Maeda 1974), and vestibulospinal neurons can project both to the cervical and to the lumb ar spinal
cords via axon collaterals (Abzug et al. 1974).
Vestibulothalamic neurons also give off descending
axons to the spinal cord (Isu et al. 1991). Thus, the
complex pattern of connections contains both converging and diverging pathways. Diversity, instead of
homogeneity, is a characteristic feature of vestibulospinal axons. Their pathway, which is composed of
multiple anatomical subunits, may selectively coordinate the activity of combinations of interneurons
and motor neurons (Rose et al. 1996).
Tonic otolithic reflexes respond to changes in linear acceleration largely via the lateral vestibulospinal
tract (Fig. 1.5). They produce excitation of the ipsilateral extensor motor neurons of the limb and inhibition of reciprocal flexor motor neurons (Wilson
and Peterson 1978; Markharn 1987). The reticulospinal tract originates in the bulbar reticular formation (Peterson 1984); most pontomedullary
reticulospinal neurons receive otolithic or combined
otolith-canal input and project mainly to the lumb ar
and less to the cervical cord (Bolton et al. 1992). In
response to stimuli in vertical planes, pontomedullary reticulospinal fibres are best suited to
contribute to otolith reflexes (Wilson 1993).
Thus, three major pathways convey vestibular
information to the spinal cord (Brodal et al. 1962):

the lateral vestibulospinal tract from the lateral


vestibular nuc!eus of Deiter to the lumbosacral
spinal cord,
the medial vestibulospinal tract from the medial
vestibular nuc!eus to the cervicothoracic spinal
cord, and
the reticulospinal tract from the pontomedullary
reticulospinal neurons to the lumb ar cord.

The spinal course and the location of human


vestibulospinal, reticulospinal, and descending propriospinal nerve fibres are characterised by controversy and vagueness in the literature (Nathan et al.
1996). Most studies concentrate on the lateral and
medial vestibulospinal tracts and on their action on
the spinal motor neurons; however, both the rostral
and the caudal areas of the vestibular nuc!ei give rise
to axons that are located in the dorsal and dorsolateral funiculi and terminate in the dorsal horn (Isu et
al. 1996).
Vestibulospinal reflexes originating in the maculae of otoliths have been established experimentally
both in humans and in animals exposed to sudden
unexpected falls. Under such conditions, EMG

MLF
MVST

cervical cord

lhotacal cord
lumbal cord

ololithic organa
(semiclrcular canala)

..... ....... --.--"


~.

~f~r ' I r
). '

mOlor neuron

Fig.l.5. Schematic representation of vestibulospinal pathways


which link the labyrinthine receptors with the antigravity motor
neurons in the cervical and lumbar spinal cord to maintain head
and body posture. Linear or angular head acceleration leads to
ipsilaterally increased tone in extensor muscles (open eircles) and
reciprocal decreased tone in flexor muscles (filled eircles) via the
descending lateral vestibulospinal tract (LVST). The dashed lines
represent probable but not yet established connections (after
Markham 1987). The medial vestibulospinal tract (MVST) travels
through the medial longitudinal fasciculus (MLF) subserving
mainly the vestibulocollic reflex for stabilisation of the head in
space.

responses were recorded in leg extensor musc!es


some 60-80 ms after on set of free fall (Melvill Iones
and Watt 1971). Even faster otolith reflex effects have
been described in lower leg musc!es in humans, with
latencies shorter than 45 ms (Fries et al. 1988). Thus,
otolithic input plays an important role in rapid
adjustment of postural balance in addition to the
short latency stretch reflexes mediated somatosensorily (Dietz et al. 1987) and long-Ioop mechanisms
(Diener et al. 1984). The coactivation pattern of

Vertigo

12

extensor and flexor leg musc1es does not fit the


traditional view of the crossed flexor/extensoractivation/inhibition relationship obtained in decerebrated cat preparations (Roberts 1978). It confirms,
however, observations on postural reactions following toe-up perturbations during upright stance,
which were interpreted to be vestibulospinal reflexes
(Allum and Pfaltz 1985). Horstmann and Dietz
(1988) reported that the early EMG responses in leg
musc1es, which are evoked through head movements, are absent in patients without vestibular
function; this finding supports the existence of fast
vestibulospinal reflexes. However, the deficit in
vestibulospinal input accounted for only 10% reduction of leg musc1e response amplitude after linear
displacement. When subjected to platform tilt
around the ankle joint and with toes-up, patients
with bilateral vestibular deficits exhibited a decrease
of long latency activity in the tibialis anterior and
soleus musc1es (Diener and Dichgans 1988; Allum et
aI. 1988).
Direction-specific postural deviation, for example
in Romberg or Unterberger testing, helps to identify
the side of the lesion in patients, but one should bear
in mind that vestibulospinal reflexes do not trigger
rigid patterns of activated musc1es. The set of
musc1es activated in response to a particular stimulus varies with varying body posture (see vestibulospinal responses in otolithic Tullio phenomenon,
p. 108), balance strategies, and even with viewing
conditions (Paulus et aI. 1984). This flexibility is
necessary to maintain balance in situations involving combined voluntary (active) and involuntary
(passive) stimulation, for example when walking on
a rolling ship.
Although fast vestibulospinal reflexes exist, their
functional significance seems limited, and the tonic
labyrinthine reflexes as demonstrated in the decerebrate animal (Sherrington 1906; Magnus 1924) are
largely inhibited in the mature animal and human.
Control of head and body position in space is not
dominated by vestibulospinal reflexes; they are part
of a multisensory sensorimotor system. This complex neuronal network involves the cerebellum, the
cerebral cortex, the basal ganglia, and the spinal cord:

The vestibulospinal reflexes are controlled by the


spinocerebellum, whose most important part
constitutes the anterior lobe of the cerebellum
(this region receives spinocerebellar afferents
and projects back to the spinal cord via the reticulospinal tract and other pathways), and the
vestibulo-cerebellum incorporating the lower
vermis and flocculus, which has reciprocal connections with the vestibular nuc1ei and the pon-

tine reticular formation (BrodaI1981; Fetter and


Dichgans 1996). The cerebellar anterior vermis
exerts a positive influence on the basic gain of
vestibulospinal reflexes (Andre et aI. 1994;
Manzoni et aI. 1997).
Vestibular cortex areas project to the brainstem
vestibular nuc1ei (Akbarian et aI. 1993), and thus '
can modulate not only the vestibulo-ocular but
also vestibulospinal reflexes. Furthermore, voluntary changes in position and locomotion must
hierarchically control brainstem reflexes, which
might interfere with the intended pro gram. The
cortex and basal ganglia are involved in this
controI.
There must be a c10se interaction between leg,
trunk, neck proprioception and vestibulospinal
reflexes for maintenance of balance. Using combinations of support-surface rotation and rearward translation, Allum et aI. (1995) showed the
confluence of knee and trunk proprioceptive and
vestibulospinal inputs, rather than either alone
in isolation, which provides the musc1e synergy
necessary for normal balance corrections.
Mergner et aI. (1997) made c1ear that if vestibular
input is to be used for postural control of upright
stance, this necessarily involves co ordinate transformations. They developed a model of how a
proprioceptive feedback loop for body stabilisation relative to the support surface could co operate with one for body stabilisation in space. One
precondition is back-and-forth channelling of
information for the integration of vestibular and
proprioceptive input in the spinal cord. The
vestibular system or the visual scene can be used
as a reference. The latter requires back-and-forth
channelling (forth, to evaluate the kinematic
state of the visual reference; back, because all
other perceptions then have to be referred to this
object) and a mechanism for deciding which reference to use (Mergner et aI. 1997).

The integration of vestibulospinal reflexes into cerebellar, spinal cord, reticular formation, basal ganglia,
and cortical mechanisms makes it easier to understand that the pattern of musc1e responses to the
same vestibular stimuli changes with changes in
body posture or voluntary movements and visual
conditions (see varying vestibulospinal reflex
responses, Figs. 6.6-6.12, p. 108). The finding that
chronic head tiIt produced by hemilabyrinthectomy
does not depend on direct vestibulospinal tracts
(Fukushima et al. 1988) also demonstrates the functional significance of brainstem integration centres
(p. 184), which make use of oligosynaptic fast reflexes
without exc1usively relying on them.

Introduction

13

Vestibular falls
The following conditions may give rise to symptomatic falls: cardiovascular cerebral hypoxia, epilepsy,
intoxication, ataxia, movement disorders, paresis, or
severe sensory loss. Vestibular dysfunction is, however, a significant differential diagnosis for patients
presenting with unpreventable or unexpected falls.
This is not adequately recognised by clinicians outside the field of neuro-otology.
Peripheral and central vestibular pathways run
from the labyrinths via vestibular and ocular motor

nuclei to the thalamus and vestibular cortex. Either


peripheral or central vestibular disorders cause postural instability with preferred directions of falling
(Fig. 1.6). The particular pathological mechanisms
that provoke postural instability and cause vestibular falls differ considerably, because they may result
from changes in otolith or in horizontal or vertical
semicircular canal function.
Vestibular falls may be attributed to either the
particular plane of the affected semicircular canal
or a central pathway that mediates the threedimensional vestibulo-ocular reflex in yaw, pitch,
and roll. Ipsiversive falls occur in vestibular neuritis

PREFERRED DIRECTION OF FALL

DISORDER

lateral

diagonal

otolith Tullio phenomenon


lateropu Ision (grade 1-11)
vestibular epi lepsy

fore-aft

vertieal

vestibular drop attaeks

vestibular neuritis
lateropulsion (grade III-IV)
ocular tilt reaetion
thalamie astasia

benign paroxysmal positioning vertigo


bilateral vest ibulopathy
down beat nystagmus/vertigo
upbeat nystagmus/vertigo

Fig.1.6. Schematic drawing of the directions of vestibular falls in different central and peripheral vestibular syndromes. (From
Brandt and Dieterich 1996.)

14

Vertigo

or in Wallenberg's syndrome - where they are


known as lateropulsion. Contraversive falls are
typical for the otolith Tullio phenomenon, vestibular epilepsy, and thalamic astasia. Fore-aft instability
is predominantly observed in bilateral vestibulopathy, benign paroxysmal positioning vertigo, as
weIl as in downbeat or upbeat nystagmus syndrome.
Falls can be diagonally forward (or backward) and
toward or away from the side of the lesion, depending on the site of the lesion (the ocular tiIt reaction
is ipsiversive in peripheral vestibular and medullary
lesions, but contraversive in mesencephalic lesions)
and on whether vestibular structures are excited or
inhibited.
In the following, examples will be given of peripheral and central vestibular falls with particular
emphasis on our current knowledge of how and why
patients fall (Tables 1.3 and 1.4; Brandt and Dieterich
1993).

Table 1.4. Central vestibularfalls


Disorder

Direction

Mechanism

Vestibular epilepsy

Contraversive (?)

Simple or complex
partial seizures due to
epileptic discharges of
vestibular cortex

Thalamic astasia

Contraversive
or ipsiversive

Vestibular tone
imbalance (yaw, roll ?)
in posterolateral
vestibular thalamic
lesions

Ocular tilt reaction

Contraversive in
Tone imbalance ofVOR
pontomesencephalic in roll due to lesions of
lesions
otolith and vertical
canal pathways
Ipsiversive in
pontomedullary
lesions

Paroxysmal ocular tilt


reaction

Contraversive with
peripheral vestibular
stimulation

Table 1.3. Peripheral vestibular falls


Disorder

Direction

Mechanism

Vestibular neuritis

Lateral
ipsiversive

Vestibular tone
imbalance (yaw, roll)
due to horizontal
and anterior
semicircular
canal paresis

Benign paroxysmal
positioning vertigo
(BPPV)

Forward
ipsiversive

Ampullofugal
stimulation of
posterior canal by
canalolithiasis and a
heavy clot-induced
endolymph flow

Meniere's drop attacks


Vertical
(Tumarkin's otolithic crisis) (down?)

Loss of postural tone due


to abnormal otolith
stimulation in sudden
endolymphatic fluid
pressure changes

Otolith Tullio phenomenon Backward


contraversive

Sound-induced
mechanical stimulation
of utricle by luxated
stapes footplate
(diagonal)

Vestibular paroxysmia

Forward
Neurovascular crosscontraversive (?) compression
causing ephaptic
stimulation of
vestibular nerve
(multidirectional ?)

Bilateral vestibulopathy

Multidirectional Impaired postural reflexes


fore-aft
particularly in darkness

From Brandt and Dieterich (1993).

Ipsiversive with
Pathological excitation
pontomesencephalic of otolith and vertical
stimulation
canal pathways
mediating VOR in roll

Lateropulsion
(Wallenberg's
syndrome)

Ipsiversive
diagonal

Lesion-induced tone
imbalance ofVOR in
roll and yaw with
concurrent deviation
of subjective vertical

Downbeat nystagmus/
vertigo

Backward

Lesional vestibular tone


imbalance of
VOR in pitch

From Brandt and Dieterich (1993).

Peripheral vestibular falls


Vestibular neuritis: contraversive rotational vertigo
with ipsiversive falls

In vestibular neuritis - due to a vestibular tone


imbalance - the fast phase of the spontaneous rotational nystagmus and the initial perception of
apparent body motion are directed away from the
side of the lesion, and the postural reactions initiated by vestibulospinal reflexes are usually in a direction opposite to the direction of vertigo. These result
both in the Romberg fall and in pastpointing toward
the side of the lesion. There are two sensations,
opposite in direction, and the patient may be
describing either one (Fig. 4.1). The first is the purely
subjective sense of self-motion in the direction of
the nystagmus fast phases, which is not associated
with any measurable body sway. The second is the
compensatory vestibulospinal re action resulting in

Introduction
objective, measurable destabilisation in the direction
opposite to the fast phases (Brandt and Daroff 1980).
Benign paroxysmal positioning vertigo (BPPV):
forward falls produced by canalolithiasis of the
posterior semicircular canal

Posturographic measurements in patients with


BPPV, in whom attacks were elicited by head tiIt
while standing on a force-measuring platform
(Bchele and Brandt 1979) revealed a characteristic
pattern of postural instability. After a short latency
patients exhibit large sway amplitudes, predominantly in the fore-aft direction (Figs. 16.4-16.6) with
a mean sway frequency range < 3 Hz. Instability
decreases over 10 to 30 sparallel to the reduction of
nystagmus and the sensation of vertigo. When subjects dose their eyes, the acute destabilisation may
lead to an almost irresistible tendency to fall.
Posturographic data show a shift of the mean position of the cent re of gravity forward and toward the
direction of the head tilt (Figs. 16.4, 16.6). The measurable shift of the cent re of gravity in the forward
direction and ipsiversive to the tiIted head can be
interpreted as the motor compensation for the initial
subjective vertigo in the opposite direction, the diagonal plane corresponding to the spatial plane and
working range of the ipsilateral posterior canal
(Bchele and Brandt 1986).
Meniere's drop attacks (Tumarkin's otolithic crisis)

In Meniere's disease periodic endolymphatic membrane ruptures with subsequent transient potassium
excitation and palsy of vestibular nerve fibres cause
vertigo attacks and postural instability with characteristics similar to those in vestibular neuritis. The
direction of nystagmus and vertigo changes during
the attack (p. 89) and also depends on the location of
the membranous leakage in relation to either the
posterior or horizontal ampullary nerve. Rarely,
vestibular drop attacks (Tumarkin's otolithic crisis;
Tumarkin 1936) occur in early and late stages of
endolymphatic hydrops (Baloh et al. 1990) when
sudden changes in endolymphatic fluid pressure
cause non-physiological end-organ stimulation
(deformation of utride or saccule membrane?) with
a reflex-like vestibulospinal loss of postural tone.
Patients fall without warning; they remain conscious
but lose voluntary control of balance. Sometimes
during a vestibular drop attack patients have the
feeling that they are being pushed or thrown to the
ground. However, slower sensations involving
apparent tilts of the surroundings also occur, possibly resuIting in forward, backward, or lateral body
tilt.

15

Otolith Tullio phenomenon: contraversive ocular tilt


reaction (OTR) and fall

Sound-induced vestibular symptoms such as vertigo,


nystagmus, oscillopsia and postural imbalance in
patients with perilymph fistulas are commonly
known as the Tullio phenomenon (Tullio 1929). An
otolith Tullio phenomenon due to hypermobile
stapes footplate typically manifests with the pattern
of sound-induced paroxysms of OTR (Dieterich et
al. 1989). The patients complain of distressing
attacks of vertical oblique and rotatory oscillopsia
(apparent tilt of the visual scene) and of falls toward
the unaffected ear and backward elicited by loud
sounds (Fig. 6.6). The cause is a non-physiological
mechanical otolith stimulation. Surgical exploration
of the middle ear may reveal a subluxated stapes
footplate with a hypertrophic stapedius musde causing pathologically large amplitude movements during the stapedius reflex. The otolith lies directly
adjacent to the stapes footplate.
Bilateral vestibulopathy with predominant forward
and backward falls

Bilateralloss of vestibular function causes unsteadiness of gait, particularly in the dark, and - because
of the insufficiency of the vestibulo-ocular reflex at
higher frequencies - oscillopsia, associated with
head movements or when walking. These patients
complain of oscillopsia and imbalance, and the condition can be identified by the decreased ocular
motor responses to caloric irrigation and angular
acceleration (Baloh et al. 1989). Measurements of
postural instability show the largest amplitude in the
fore-aft direction (Fig. 8.6), corresponding to the
predominant direction of fall. In cases of body perturbations, falls mayaIso occur sideways, particularly
in darkness when vision cannot substitute sufficiently
for the vestibular deficit. The lack of one channel of
sensory input - important as it is for demanding balancing tasks in sport - rarely manifests as dinically
significant instability. In the absence of sensory
information from two of the stabilising systems, postural control may be severely impaired as, for exampIe, in a patient with sensory polyneuropathy and/or
with bilateral vestibulopathy (Fig. 30.3) under
restricted visual conditions (darkness).

Central vestibular falls


Vestibular epilepsy with contraversive vertigo and
falls

From the few detailed reports on the direction of


apparent self-motion and surround motion in

Vertigo

16

patients with vestibular epilepsy (p. 234), it is most


likely that the direction of perceived self-motion,
measurable body motion, and eye deviation is contraversive to the epileptic focus, whereas simultaneously perceived surroun~ motion may be
ipsiversive (Foerster 1936). Actual body movements
do not represent vestibulospinal compensations of
perceived vertigo but an epileptic response. Rotatory
seizures in rare "volvular epilepsy" (p. 234) are characterised by paroxysmal, repetitive walking in small
circles. Vestibular seizures can manifest without any
objective eye and body movements, as Foerster (1936)
described in his stimulation experiments.
Thalamic astasia with contraversive or ipsiversive
falls?

There are a few instances of presumed central


vestibular dysfunction in which patients without
paresis or sensory or cerebellar deficits are unable to
maintain an unsupported, upright posture. The conditions are thalamic astasia (p. 192), lateropulsion in
Wallenberg's syndrome (p. 309), and OTR (p. 179).
Postural imbalance with a transient tendency to fall
has been noted following therapeutic thalamotomy
and thalamic haemorrhages (Verma and Maheshwari
1986). Thalamic astasia, as described by Masdeu and
Gorelick (1988), occurred as a resuIt of lesions with
different causes, all primarily involving superoposterolateral portions of the thalamus but sparing
the rubral region. It is our own experience in some
30 patients with thalamic infarctions that the posterolateral type may cause both contraversive or ipsiversive postural instability (Dieterich and Brandt 1995).
Ocular tift reaction: ipsiversive in caudal,
contraversive in upper brainstem lesions

Ocular tilt reaction (OTR) is a vestibular tone imbalance involving the vertical vestibulo-ocular reflex in
the roll plane (p. 180). It represents a fundamental
pattern of coordinated eye-head-roll motion and
body tilt, is based on both otolith and vertical canal
input, and is mediated by the graviceptive pathways
from the labyrinths via the rostral medial and
superior vestibular nuclei and the contralateral
medial longitudinal fascicle to the rostral midbrain
tegmentum. The OTR consists of lateral head tiIt,
skew deviation of the eyes (hypotropia of the undermost eye), and ocular torsion (clockwise with head
tiIt left; counterclockwise with head tiIt right). It was
first clearly delineated during electrical stimulation
of the interstitial nucleus of Cajal (Westheimer and
Blair 1975).
OTR is not a rare condition. In acute unilateral
brainstem infarctions it can be detected in about

20% of cases if a careful examination for ocular torsion of the eyes (fundus photographs), subtle skew
deviation, and subjective visual vertical (Brandt and
Dieterich 1992) is carried out. OTR and concurrent
body tiIt are always ipsiversive in pontomedullary
lesions (Brandt and Dieterich 1987), whereas OTR
and concurrent body tiIt are always contraversive in
pontomesencephalic lesions (Halmagyi et al. 1990).
Lateropulsion in Wallenberg's syndrome: ipsiversive
falls and adjustments of perceived vertical

Lateropulsion ofthe body (Fig.19.3) is a well-known


transient feature of lateral medullary infarction in
which patients cannot prevent ipsiversive lateral
falls. We believe that subjective vertigo is usually
absent in these patients because there is no sensory
mismatch. The lesion causes a deviation of the perceived vertical. Individual muItisensory regulation of
posture is then adjusted not to the true vertical but
to the pathologically deviated internal representation of verticality produced by the lesion. The more
pronounced the lateropulsion, the greater the deviations of subjective visual vertical adjustments (Fig.
19.4). Thus, these patients fall without realising that
it is their active shift of the cent re of gravity (lateropulsion) that causes the imbalance. Here also it is
the incorrect central computation of verticality
(despite correct peripheral sensory signals from the
otoliths) that is responsible for postural imbalance
(Dieterich and Brandt 1992).
Downbeat nystagmus syndrome with backward falls

Downbeat nystagmus in the primary gaze position


(Cogan 1968), or in particular on lateral gaze, is
often accompanied by oscillopsia and postural instability. Posturographic measurements show a typical
postural imbalance with a striking fore-aft body
sway (Bchele et al. 1983) and a tendency to fall
backward (Figs. 1.7; 11.1-11.3). This fore-aft postural
instability can be interpreted as a direction-specific
tone imbalance of the VOR in pitch, due either to a
lesion in the floor of the fourth ventricle or to a
bilateral lesion of the flocculus. Thus, downbeat
nystagmus is not simply an oculomotor syndrome
but a central vestibular syndrome comprising oculomotor, postural, and perceptual effects.

Vestibular autonomie regulation


Vestibular autonomic regulation is known to play an
important role in the different forms of motion sick-

Introduetion

17

EYES OPEN

EYES CLOSED

normal

IOmm
I---i

autonomie functions are significantly modulated by


vestibular input:

I-------l

Vestibular Neuritis

..... ,

.'

S.E. cl 54

I-------l

Downbeat Nystagmus Syndrome

K.F. d' 48

Fig. 1.7. Postural instability in downbeat nystagmus.


Histograms for fore-aft (A-P) and lateral (R,L) postural sway during upright stance with eyes open (/eft) and eyes closed (fight)
obtained with a force-measuring platform. For comparison, see
registrations of body sway in anormal subject (top) and in a
patient with vestibular neuritis (middle). Preferred direction of
postural instability and body sway is in the fore-aft direction.
(From Brandt and Dieterich 1993.)

ness (see Chap. 33). The pleasurable effeets of


vestibular stimuli on the autonomie system are perhaps even more widely known, to judge from the
droves of people eonverging on amusement park
rides. Also young mothers quiekly diseover that by
gently rocking the baby earriage they can transform
an uncomfortable whiner into a contented baby, who
sometimes even falls asleep.
Vestibular nuclei retieular formation and midline
vestibulo-cerebellar structures integrate mainly
vestibular and autonomie activities. The following

susceptibility to motion sickness (p. 490) and


spaee sickness (Oman 1998; Parker 1998) dependent on preserved function of the vestibular system,
respiratory control by adjustments of the activity
of respiratory muscles during changes in body
position (Yates and Miller 1998),
eardiovaseular control during orthostatic stress
or ehanges in gravitational forees (Biaggioni et
al. 1998; Cui et al. 1997; Convertino 1998),
control of sleep, in partieular REM sleep and
dreaming (Hobson et al. 1998),
produetion of affeetive and emotional responses,
possibly linked to anxiety dis orders (p. 459),
including agoraphobia and panie attacks
(Furman et al. 1998),
modulation of perceived verticality (Mittelstaedt
1996).

The involvement of vestibulo-autonomic structures


in the mechanism of nausea and vomiting is treated
in detail in Chap. 33 (Motion sickness; p. 485) and
Chap. 18 (Paroxysmal positioning vomiting; p. 293).
Changes in body posture, e.g. from supine to standing, eause orthostatic hypotension due to blood
pooling in the lower limbs and influence the respiratory muscles to temporarily impair respiration. To
effeetively maintain homeostasis during ehanges
of body posture, the compensation of the effects
movement has on eirculation and respiration must
even begin before the internal environment has been
affected. The vestibular system might aeeomplish
this by detecting head movement and position and
thus provide "feed-forward information" to the autonomie centres (Yates et al. 1998). The vestibular
system has access to respiratory muscles by whieh it
adjusts the aetivity of the multifunetional respiratory
muscles necessary to offset the meehanieal eonstraints on these muscles which oecur during
changes ofbody position (Yates and Miller 1998).
Animal experiments have shown that the vestibular system contributes to orthostatie cardiovascular
reflexes in the cat (Doba and Reis 1974) and monkey
(Satake et al. 1991). Likewise, vestibular input plays a
role in cardiovascular control in humans (Biaggioni
et al. 1998). In particular, experiments have shown
that semicircular canal stimulation interaets with
carotid baroreceptor reflex control of heart rate
(Convertino 1998).
Sleep and dreams are also influenced by the
vestibular component. Dreaming subjects perceive
themselves as constantly moving through the dream
space; such features are based on REM sleep (Porte

Vertigo

18

and Hobson 1996). The "vestibular eharaeter" of


these sensations include flying, floating, swimming,
spinning, twitehing, or turning, whieh dreamers generally eonsider exciting or pleasurable (MeCarley
and Hoffman 1981; Porte and Hobson 1996).
"Vestibular dreaming" is facilitated by sleeping in a
swinging hammoek (Leslie and Ogilvie 1996). The
vestibular nuclei are an integral part of the pontomedullary retieular neuronal network whieh eontrols the sleep-wake eycle (Lorente de No 1933;
Pompeiano et al. 1991). Drawing on Moruzzi's
original idea that sleep subserves neuronal plastieity,
Hobson and eo-workers (1998) have hypothesised
that vestibular adaptation simultaneously provokes
and is enhaneed by REM sleep under both normal
gravity and spaee-flight eonditions.
Finally, several reports have linked vestibular
funetion to panie disorder and agoraphobia (Chaps.
31 and 32; p.456). It has been speeulated that
aseending vestibulo-autonomie connections form a
basis for an association between vestibular funetion
and panie attaeks (Furman et al. 1998). These
authors stress that vestibular-indueed autonomie
symptoms include nausea, malaise, drowsiness,
abdominal awareness, anxiety, distress, dread and
redueed vigilanee. Vestibular-indueed autonomie
signs include ehanges in salivation, gastrie motility,
vomiting, endoerine responses, "eoldness", sweating,
pallor, inereased blood flow to skeletal muscles and
ehanges in he art rate.
The examples given so far deal with autonomie
effeets that are indueed or modulated by vestibular
stimuli or vestibular dysfunetion. Another hitherto
poorly understood vestibulo-autonomie interaction
might be relevant: the eonvergenee of vestibular and
visceral autonomie input for the internal representation of vertieality (Mittelstaedt 1996). A specifie, as
yet unidentified abdominal organ would not be
required for the pereeption of gravity; use eould be
made of viseeral proprioeeption, the distribution of
the abdominal organs, or blood pooling.

Neuroanatomie substrates
Neuroanatomie substrates for vestibulo-autonomie
interactions are based on the eonvergenee of
vestibular and autonomie pathways, partieularly in
the vestibular nuclei and the arehieerebellum. As
reviewed by Balaban and Porter (1998),

"a eaudal region (eaudal medial vestibular nucleus


and the inferior vestibular nucleus) eontributes
both (a) light deseending projeetions to the
nucleus of the solitary traet, the dorsal motor
vagal nucleus, the nucleus ambiguus, the ventro-

lateral medullary retieular formation, the nucleus


raphe magnus, and the lateral medullary
tegmentum, and (b) aseending projeetions to the
parabrachial nucleus.
A rostral region (the superior vestibular nucleus
and the rostral pole of the medial vestibular
nucleus) though, eontributes only aseending projeetions to the parabrachial region."

It is assumed that

the deseending pathways are involved in


medullary eardiovaseular and respiratory eontrol and in the generation of motion siekness
with nausea and vomiting.
The aseending pathways that reaeh the hypothalamus, amygdaloid nucleus, the infralimbie and
insular cortex eontribute to affeetive and emotional responses.

There is evidenee that vestibulo-autonomie pathways are under inhibitory eerebellar eontrol as is the
vestibulo-oeular reflex (by the floeeulonodular lobe)
and the vestibulospinal pathways (by the anterior
lobe). Balaban and Porter (1998) delineate "four
medial eerebellar regions that appear to influenee
vestibulo-autonomie funetion:
an intermediolateral site on the border oflobula
IX and the nodulus (lateral nodulus-uvula
region),
2. a eaudal, posterior lobe region in zone A of lobula
IX (medial uvular region),
3. a rostral, posterior lobe region in zone A of
lobules VIIa through VIlla (rostral posterior lobe
region), and
4. an anterior lobe region within zone A of lobules
I-Ill."
1.

It is still not possible to clearly eoneeptualise the

strueture and funetion of the largely unknown eentral autonomie network. The main barrier eontinues
to be its eomplex integration in organie and psyehie
processes that operate during internal body events
and external sensory stimulation. It is mueh easier to
establish a eoneept for sensorimotor eontrol of eye
movements with defined input/output relations.
Future studies on the vestibular system, however,
should pay heed to the autonomie responses inherent, for example, in stimulation by virtual motion
displays or ealorie or galvanie stimulation. The latter
is of partieular importanee for interpreting eerebral
aetivation patterns in fMRI or PET studies. Many of
the aetivated areas will be found to be related to
autonomie rather than simply pereeptual or oeular
motor funetions. Autonomie signs and symptoms

Introduction
should also be recognised as part of peripheral or
central vestibular disorders such as vertigo attaeks
in Meniere's disease or basilar migraine. Management of vertigo should also try to reduee distressing
autonomie symptoms.

References
Abzug C, Maeda M, Peterson BW, Wilson VJ (1974) Cervical
branching of lumb ar vestibulo-spinal axons. J Physiol (London)
243:499-522
Akbarian S, Grsser O-J, Guldin WO (1993) Corticofugal projections to the vestibular nuclei in squirrel monkeys: further evidence of multiple cortical vestibular fields. J Comp Neurol
332:89-104
Allum JHJ, Pfaltz CR (1985) Visual and vestibular contributions to
pitch-sway stabilisation in the ankle muscles of normals and
patients with bilateral peripheral vestibular defieits. Exp Brain
Res 58:82-94
Allum JHJ, Keshner EA, Honegger F, Pfaltz CR (1988) Indicators of
the influence a peripheral vestibular deficit has on vestibulospinal reflex responses controlling postural stability. Acta
Otolaryngol (Stockh) 106:252-263
Allum JH, Honegger F,Acuna H (1995) Differential control ofleg
and trunk muscle activity by vestibulo-spinal and proprioceptive signals during human balance corrections. Acta
Otolaryngol (Stockh) 115:124-129
Anastasio TJ (1992) Simulating vestibular compensation using
recurrent back-propagation. Biol Cybern 66:389-397
Anastasio TI, Robinson DA (1989) Distributed parallel processing
in the vestibulo-oculomotor system. Neural Computation
1:230-241
Anastasio TJ, Robinson DA (1990) Distributed parallel processing
in the vertical vestibulo-oculomotor reflex: learning networks
compared to tensor theory. Biol Cybern 63:161-167
Andre P, d' Ascanio P, Manzoni D, Pompeiano 0 (1994) Depression
of the vestibulospinal reflex by intravermal microinjection of
GABA A and GABA B agonists in the decerebrate cat. J Vestib Res
4:251-268
Arnold DB, Robinson DA (1991) A learning network of the neural
integrator ofthe oculomotor system. Biol Cybern 64:447-454
Assaiante C, Amblard B (1990) Head stabilisation in space while
walking: effect of visual deprivation in children and adults. In:
Brandt T, Paulus W, Bles W, Dieterich M, Krafczyk S, Straube A
(eds) Disorders of posture and gait. Thieme, Stuttgart, pp
229-232
Baker J, Goldberg I, Hermann G, Peterson B (1984) Optimal
response planes and canal convergence in secondary neurons
in vestibular nuclei of alert cats. Brain Res 294:133-137
Balaban CD, Porter JD (1998) Neuroanatomie substrates for
vestibulo-autonomic interactions. J Vestib Res 8:7-16
Baloh RW, Beykirch K, Honrubia V, Yee RD (1988) Eye movements
induced by linear acceleration on parallel swing. J Neurophysiol
60:2000-2013
Baloh RW, Jacobson K, Honrubia V (1989) Idiopathic bilateral
vestibulopathy. Neurology 39:272-275
Baloh RW, Jacobson K, Winder T (1990) Drop attacks with
Meniere's syndrome. Ann NeuroI28:384-387
Barret HJ, Hood JD (1988) Transfer of optokinetic activity to
vestibular nystagmus. Acta Otolaryngol (Stockh) 105:318-327
Biaggioni I, Costa F, Kaufmann H (1998) Vestibular influence on
autonomie cardiovascular control in humans. J Vestib Res
8:35-41
Bishop PO (1978) Orientation and position disparities in stereop-

19
sis. In: Cool SJ, Smith EL (eds) Frontiers in visual science.
Springer, New York, pp 336-350
Bolton PS, Goto T, Schor RH, Wilson VJ, Yamagata Y, Yates BJ
(1992) Response of pontomedullary reticulospinal neurons to
vestibular stimuli in vertical planes: role in vertical vestibulospinal reflexes of the decerebrate cat. J Neurophysiol
67:639-647
Brandt Th, Daroff RB (1980) The multisensory physiological and
pathological vertigo syndromes.Ann NeuroI7:195-203
Brandt T, Dieterich M (1987) Pathological eye-head co ordination
in roll: tonic ocular tilt reaction in mesencephalic and
medullary lesions. Brain 110:649-666
Brandt T, Dieterich M (1992) Cyclorotation of the eyes and subjective visual vertical in acute vascular (vestibular) brainstem
lesions.Ann NY Acad Sei 658:537-549
Brandt T, Dieterich M (1993) Vestibular falls. J Vestib Res 3:3-14
Brandt T, Dieterich M (1996) Postural imbalance in peripheral
and central vestibular disorders. In: Bronstein A, Brandt T,
Woollacott M (eds) Clinical disorders of balance, posture and
gait. Edward Arnold, London, pp 131-146
Brodal A (1981) Neurological anatomy. Oxford University Press,
NewYork
Brodal A, Pompeiano 0, Walberg F (1962) The vestibular nuclei
and their connections, anatomy and functional correlations.
Oliver & Boyd, Edinburgh
Bchele W, Brandt Th (1979) Vestibulo-spinal ataxia in benign
paroxsymal positional vertigo. Agressologie 20:221-222
Bchele W, Brandt Th (1986) Benign paroxysmal positional vertigo and posture. In: Bles W, Brandt T (eds) Disorders of posture
and gait. Elsevier,Amsterdam, pp 141-156
Bchele W, Brandt Th, Degner D (1983) Ataxia and oseillopsia in
downbeat nystagmus/vertigo syndrome. Adv Oto-RhinoLaryngoI30:291-297
Cannon SC, Robinson DA (1985) An improved neural-network
model for the neural integrator of the oculomotor system: more
realistic neuron behavior. Biol Cybern 53:93-108
Cannon SC, Robinson DA (1987) Loss of the neural integrator of
the oculomotor system from brain stern lesions in monkey. J
NeurophysioI57:1383-1409
Cogan DG (1968) Downbeat nystagmus. Arch Ophthalmol
80:757-768
Cohen B, Henn V (1988) Representation of three-dimensional
space in the vestibular, oculomotor, and visual systems. Ann NY
Acad Sei 545
Collewijn H (1989) The vestibulo-ocular reflex: is it an independent subsystem? Rev Neurol (Paris) 145:502-512
Collewijn H, van der Steen J, Ferman L, Jansen TC (1985) Human
ocular counterroll: assessment of static and dynamic properties
from electromagnetic scleral coi! recordings. Exp Brain Res
59:185-196
Convertino VA (1998) Interaction of semieircular canal stimulation with carotid baroreceptor reflex control of heart rate. J
Vestib Res 8:43-49
Crawford JD, Cadera W, Vilis T (1991) Generation of torsional and
vertical eye position signals by the interstitial nucleus of Cajal.
Science 252:1551-1553
Cui I, Mukai C, Iwase S, Sawasaki N, Kitazawa H, Mano T,
Sugiyama Y, Wada Y (1997) Response to vestibular stimulation
of sympathetic outflow to muscle in humans. J Auton Nerv Sys
66:154-162
Darlot C, Denise P, Droulez J, Cohen B, Berthoz A (1988) Eye
movements induced by off-vertical axis rotation (OVAR) at
small angles oftilt. Exp Brain Res 73:91-105
Diener H C, Dichgans J (1988) On the role of vestibular, visual, and
somatosensory information for dynamic postural control in
humans. Prog Brain Res 76:253-262
Diener HC, Dichgans J, Guschlbauer B, Mau H (1984) The significance of proprioception on postural stabilization as assessed by
ischemia. Brain Res 296:103-109

20
Dieterich M, Brandt T (1992) Wallenberg's syndrome: lateropulsion, cyclorotation and subjective visual vertical in thirty-six
patients. Ann NeuroI31:399-408
Dieterich M, Brandt T (1995) Vestibulo-ocular reflex. Curr Opin
Neurol 8:83-88
Dieterich M, Brandt Th, Fries W (1989) Otolith function in man:
results from a case of otolith Tullio phenomenon. Brain
112:1377-1392
Dietz V, Quintern J, Sillem M ( 1987) Stumbling reactions in man:
Significance of proprioceptive and pre-programmed mechanism. J Physiol (London) 386:140-163
Doba N, Reis DJ (1974) Role of cerebellum and vestibular apparatus in regulation of orthostatic reflexes in the cat. Circ Res
34:9-18
Ferman L, Collewijn H, Jansen TC, Van den BergV (1987) Human
gaze stability in the horizontal, vertical and torsional direction
during voluntary head movements, evaluated with a three
dimensional scleral induction coil technique. Vision Res
27:811-828
Fetter M, Dichgans J (1996) How do the vestibulo-spinal reflexes
work? In: Baloh RW, Halmagyi GM (eds) Disorders of the
vestibular system. Oxford University Press, New York, Oxford,
pp 105-112
Foerster 0 (1936) Sensible corticale Felder. In: Bumke 0, Foerster
o (eds) Handbuch der Neurologie, vol VI. Springer, Berlin, pp
358-448
Fries W, Dieterich M, Brandt Th (1988) Otolithic control of posture: Vestibulo-spinal reflexes in a patient with a Tullio
phenomenon. Adv OtorhinolaryngoI41:162-165
Fukushima K, Fukushima J, Kato M (1988) Head tilt produced by
hemilabyrinthectomy does not depend on the direct vestibulospinal tracts. Brain Behav EvoI32:181-186
Furman JM, Jacob RG, Redfern MS (1998) Clinical evidence that
the vestibular system participates in autonomic control. J Vestib
Res 8:27-34
GrafW, Simpson JI (1981) The relations between the semicircular
canals, the optic axis, and the extraocular muscles in lateraleyed and frontal-eyed animals. In: Fuchs AF, Becker W (eds)
Progress in oculomotor research. Elsevier, Amsterdam,
pp. 409-417
GrafW, Brunken WJ (1984) Elasmobranch oculomotor organization: anatomical and theoretical aspects of the phylogenetic
development of vestibulo-oculomotor connectivity. J Comp
NeuroI227:569-581
GrafW, Ezure K (1986) Morphology of vertical canal related second order vestibular neurons in the cat. Exp Brain Res 63:35-48
Graf W, McCrea RA, Baker R (1983) Morphology of posterior
canal-related secondary vestibular neurons in rabbit and cat.
Exp Brain Res 52:125-138
Gresty MA, Bronstein AM (1992) Testing otolithic function. Br J
AudioI26:125-136
Gresty MA, Bronstein AM, Brandt Th, Dieterich M (1992)
Neurology of otolithic function: peripheral and central disorders. Brain 115:647-673
Grossman GE, Leigh RJ (1990) Instability of gaze during locomotion in patients with deficient vestibular function. Ann Neurol
27:528-532
Guedry FE (1965) Orientation of the rotation axis relative to gravity: its influence on nystagmus and the sensation of rotation.
Acta Otolaryngol (Stockh) 60:30-48
Halmagyi GM, Curthoys IS (1988) A clinical sign of canal pareses.
Arch Neurol45: 737-739
Halmagyi GM, Brandt Th, Dieterich M, Curthoys IS, Stark RJ, Hoyt
WE (1990) Tonic contraversive ocular tilt reaction due to unilateral meso-diencephalic lesion. Neurology 40: 1503-1509
Hess BJM, Dieringer N (1991) Spatial organization of linear
vestibuloocular reflexes of the rat: responses during horizontal
and verticallinear acceleration. J NeurophysioI66:1805-1818
Hobson JA, Stickgold R, Pace-Schott EF, Leslie KR (1998) Sieep

Vertigo
and vestibular adaptation: implications for function in microgravity. J Vestib Res 8:81-94
Horstmann GA, Dietz V (1988) The relative contributions of the
vestibular and muscle proprioceptive input to the stabilization
of human posture; a new experimental approach. Neurosci Lett
95:179-184
Isu N, Sakuma A, Kitahara M, Uchino Y (1991) Vestibulo-thalamic
neurons give off descending axons to the spinal cord. Acta
Otolaryngol (Stockh) SuppI481:216-220
Isu N, Thomson DB, Wilson VJ (1996) Vestibulospinal effects on
neurons in different regions of the gray matter of the cat upper
cervical cord. J Neurophysiol 76:2439-2446
Iwamoto Y, Perlmutter SI, Baker JF, Peterson BW (1996) Spatial
co ordination by descending vestibular signals. 2. Response
properties of medial and lateral vestibulospinal tract neurons
in alert and decerebrate cats. Exp Brain Res 108:85-100
KerteszAE (1983) Vertical and cyclofunctional disparityvergence.
In: Schor CS, Ciuffreda KJ (eds) Vergence eye movements: basic
and clinical aspects. Butterworths, Boston, pp 317-348
King OS, Seidman SH, Leigh RJ (1992) Control of head stability
and gaze during locomotion in normal subjects and patients
with deficient vestibular function. In: Berthoz A, Graf W, Vidal
PP (eds) Second symposium on head-neck sensory-motor
system. Oxford University Press, New York, pp 568-570
Leigh RJ, Brandt T (1993) Areevaluation of the vestibulo-ocular
reflex: new ideas of its purpose, properties, neural substrate,
and disorders. Neurology 43: 1288-1295
Leigh RJ, Zee DS (1999) Neurology of eye movements, 3rd ed. FA
Davis, Philadelphia
Leigh RJ, Maas EF, Grossman GE, Robinson DA (1989) Visual cancellation of the torsional vestibulo-ocular reflex. Exp Brain Res
75:221-226
Leigh RJ, Sawyer RN, Grant MP, Seidman SH (1992) High frequency
vestibuloocular reflex as a diagnostic tool. Ann NY Acad Sci
656:305-314
Leslie KR, Ogilvie R (1996) Vestibular dreams: the effect of rocking on dream mentation. Dreaming 6:1-16
Lorente de No R (1933) Vestibulo-ocular reflex are. Arch Neurol
Psychiatry 30:245-291
Magnus R (1924) KrpersteIlungen. Springer-Verlag, Berlin
Manzoni D, Andre P, Pompeiano 0 (1997) Changes in gain and
spatiotemporal properties of the vestibulospinal reflex after
injection of a GABA-A agonist in the cerebellar anterior vermis.
J Vestib Res 7:7-20
Markham CH (1987) Vestibular control of muscular tone and posture. Can J Neurol Sci 14:493-496
Markham CH, Yagi T (1984) Brainstem changes in vestibular compensation. Acta Otolaryngol (Stockh) SuppI406:83-86
Masdeu JC, Gorelick PB (1988) Thalamic astasia: inabili ty to stand
after unilateral thalamic lesions. Ann NeuroI23:586-603
Maurer C, Kimming H, Trefzer A, Mergner T (1997) Visual object
localization through vestibular and neck inputs. 1: Localization
with respect to space and relative to the head and trunk midsagittal planes. J Vestib Res 7:119-135
McCarley RW, Hoffman E (1981) REM sleep dreams and the activation-synthesis hypothesis. Am J Psychiatry 138:904-912
Melvill Jones G, Watt DG (1971) Muscular control oflanding from
unexpected fall in man. J PhysioI219:729-737
Mergner T, Huber W, Becker W (1997) Vestibular-neck interaction
and transformation of sensory coordinates. J Vestib Res
7:347-367
Mittelstaedt H (1996) Somatic graviception. Biol PsychoI42:53-7 4
Moruzzi G (1965) The functional significance of sleep with
particular regard to the brain mechanisms underlying consciousness. In: Eccles JS (ed) Brain and conscious experience.
Springer, New York, pp 345-388
Nathan PW, Smith M, Deacon P (1996) Vestibulospinal, reticulospinal and descending propriospinal nerve fibres in man.
Brain 119:1809-1833

Introduction
Oman CM (1998) Sensory conflict theory and space sickness: our
changing perspective. J Vestib Res 8:51-56
Ott D, Seidman SH, Leigh RJ (1992) The stability of human eye
orientation during visual fixation. Neurosei Lett 142: 183-186
Paige GD (1989) The influence of target distance on eye movement responses during verticallinear motion. Exp Brain Res
77:585-593
Parker DE (1998) The relative roles of the otolith organs and
semicircular canals in producing space motion sickness. J
Vestib Res 8:57-59
Paulus W, Straube A, Brandt T (1984) Visual stabilization of posture. Physiological stimulus characteristics and clinical aspects.
Brain 107:1143-1163
Peterson BW (1984) The reticulospinal system and its role in the
control of movement. In: Barnes CD (ed) Brainstem control of
spinal cord function. Academic Press, Orlando, pp 27-86
Pompeiano 0, Horn E, d' Ascanio PO, Horn E, d' Ascanio P (1991)
Locus coeruleus and dorsal pontine reticular influences on the
gain of vestibulospinal reflexes. Prog Brain Res 88:435-462
Porte H, Hobson JA (1996) Fictive motion in REM sleep: a test of
dream theory. J Abnorm Psycholl 05:329-335
Pozza T, Berthoz A, Lefort L (1990) Head stabilization during various locomotor tasks in humans.1. Normal subjects. Exp Brain
Res 82:97-106
Pozza T, Berthoz A, Lefort L, Vitte E (1991) Head stabilization during various locomotor tasks in humans. 11. Patients with bilateral
vestibular deficits. Exp Brain Res 85:208-217
Roberts TDM (1978) Neurophysiology of postural mechanism,
3rd ed. Butterworth, London
Rose PK, Tourond JA, Donevan AH (1996) Morphology of single
vestibulospinal collaterals in the upper cervical spinal cord of
the cat: III collaterals originating from axons in the ventral
funiculus ipsilateral to their cells of origin. J Comp Neurol
364:16-31
Satake H, Matsunami K, Miyata H (1991) The vestibuloautonomic
function viewed from cardiac responses in centrifuged monkeys. Acta Otolaryngol481:543
Schwarz UC, Busettini C, Miles FA (1989) Ocular responses to linear motion are inversely proportional to viewing distance.
Seience 245: 1394-1396
Seidman SH, Leigh RJ (1989) The human torsional vestibuloocular reflex during rotation about an earth-vertical axis. Brain
Res 504:264 - 268
Sherrington CS (1906) The integrative action of the nervous
system. Yale University Press, New York
Shinoda Y, Ohgaki T, Sugiuchi Y, Futami T (1988) Structural basis
for three-dimensional coding in the vestibulo-spinal reflex. In:
Cohen B, Henn V (eds) Representation of three-dimensional
space in the vestibular, oculomotor, and visual systems. Ann NY
Acad Sei 545:216-227

21
Shinoda Y, Sugiuchi Y, Futami T, Ando N, Yagi J (1997) Input
patterns and pathways from the six semicircular canals to
motoneurons of neck muscles. 11. The longissimus and semispinalis muscle groups. J Neurophysiol 77:1234-1258
Snyder LH, King WM (1988) Vertical vestibulo-ocular reflex in the
cat: asymmetry and adaptation. J Neurophysiol 59:279-298
Takagi A, Sando I, Takahashi H (1989) Computer-aided threedimensional reconstruction of semicircular canals and their
cristae in man. Acta Otolaryngol (Stockh) 107:362-365
Takahashi S, Fetter M, Koenig E, Dichgans J (1990) The clinical
significance of head-shaking nystagmus in the dizzy patient.
Acta Otolaryngol (Stockh) 109:8-14
Tullio P (1929) Das Ohr und die Entstehung der Sprache und
Schrift. Urban and Schwarzenberg, Munieh.
Tumarkin A (1936) The otolithic catastrophe: a new syndrome. Br
Med ]I: 175-177
Verma AK, Maheshwari MC (1986) Hyperesthetic-ataxichemiparesis in thalamic hemorrhage. Stroke 17:49-51
Viirre E, Tweed D, Milner K, Vilis T (1986) Areexamination of the
gain of the vestibuloocular reflex. J Neurophysiol 56:439-450
Wall C III, Furman JMR (1989) Nystagmus responses in a group of
normal humans during earth-horizontal axis rotation. Acta
Otolaryngol (Stockh) 108:327-335
Weissman BM, Discenna AO, Leigh RJ (1989) Maturation of the
vestibulo-ocular reflex in normal infants during the first 2
months of life. Neurology 39:534-538
Westerheimer G, Blair SM (1975) The ocular tilt reaction: a brainstern oculomotor routine. Invest OphthalmoI14:833-839
Wilson VJ (1993) Vestibulospinal reflexes and the reticular formation. Prog Brain Res 97:211-217
Wilson VJ, Maeda M (1974) Connection between semicircular
canals and neck motor neurones in the cat. J Neurophysiol
37:346-357
Wilson VI, Peterson BW (1978) Peripheral and central substrates
ofvestibulo-spinal reflexes. Physiol Rev 58:80-105
Wilson VJ, Boyle R, Fukushima K, Rose PK, Shinoda Y, Sugiuchi Y,
Uchino Y (1995) The vestibulocollic reflex. J Vestib Res
5:147-170
Wist ER, Brandt Th, Krafczyk S (1983) Oscillopsia and retinal slip.
Evidence supporting a clinical test. Brain 106:153-168
Yates BJ, Miller AD (1998) Physiological evidence that the vestibular system partieipates in autonomie and respiratory contro!. J
Vestib Res 8:17-25
Yates BJ, Sklare DA, Frey MAB (1998) Vestibular autonomie regulation: overview and conclusions of arecent workshop at the
University of Pittsburgh. J Vestib Res 8: 1-5
Zee DS (1978) Ophthalmoscopy in examination of patients with
vestibular dis orders. Ann NeuroI3:373-374

Approaching the patient

Auramo et al. 1993), their applieation in a clinical


setting is still quite limited.
Dizziness is a vexing symptom, difficult to assess
because of its purely subjective character and its
variety of sensations. The sensation of spinning or
rota tory vertigo is much more specific; if it persists,
it undoubtedly indieates acute pathology of the
labyrinth, the vestibular nerve, or the eaudal brainstern, whieh eontains the vestibular nudel.
History taking allows the early differentiation of
vertigo and disequilibrium disorders into seven categories that serve as a practical guide for differential
diagnosis:

About 50% of all patients presenting with dizziness,


vertigo, or disequilibrium in a neurological dizziness
unit will be suffering from one of the five following
common syndromes (Table 2.1):

benign paroxysmal positioning vertigo,


phobic postural vertigo,
basilar migraine,
Meniere's disease, or
vestibular neuritis.

A clinician not familiar with dizzy patients can


most effectively deepen his knowledge by acquainting himselfwith these five most frequently met and
challenging conditions of vertigo. Diagnosis and
management of vertigo syndromes always require
interdisciplinary thinking, and history taking is still
much more important than recordings of eye movements or brain imaging techniques. Although most
clinieians welcome the attempts to develop computer interview systems for use with neuro-otologieal
patients (O'Connor et a1. 1989) and expert systems
as diagnostie aids in otoneurology (Mira et al. 1990;

1.

2.
3.
4.
5.
6.
7.

dizziness and Iight-headedness


single or recurrent attacks of (rotatory) vertigo
sustained (rotatory) vertigo
positionallpositioning vertigo
oscillopsia (apparent motion of the visual scene)
vertigo associated with auditory dysfunetion
dizziness or to-and-fro vertigo with postural
imbalance.

Dizziness and light-headedness


Table 2.1. Frequeney of different vertigo syndromes in 2010 patients
seen in a neurological diuiness unit (1989- 1997)

Diagnosis

1.
2.
3.
4.
5.
6.

Benign paroxysmal positioning vertigo


Phobie postural vertigo
Central -vestibular vertigo
Basilar migraine
Meni~re's disease
Peripheral vestibulopathy
(vestibular neuritis)
7. Bilateral vestibulopathy
8. Psychogenic vertigo (without 2.)
9. Vestibular paroxysmia
10. Perilymph fistula
Unknown aetiology
Other (central vestibular syndromes without
vertigo)

Most of us have experienced presyneopal dizziness


at sometime when rapidly standing up from a
relaxed supine or seated position. It is the best
example of this eategory (Baloh 1996), wh ich
includes orthostatic hypotens ion and eardiae
arrhythmias as weIl as the hyperventilat ion syndrome and panie attaeks, metabolic hypoglyeaemia,
or drug intoxication (Table 2.2). The underlying
causes of presyncopal or syncopal dizziness are cardiac
and non-cardiac (Table 2.3 and Table 2.4). The common mechanism is decline in cardiac output or blood
pressure with subsequent diffuse cerebral ischaemia.

Frequeney

395

19.6

320
292
159

15.9
14.5

134
52

96

6.7
2.6
2.4
2.0
0.3
4.8

316

15.8

151

49

41
5

7.9
7.5

Attacks of (rotatory) vertigo, episodic vertigo


Recurrent vertigo attacks lasting for some seconds
or minutes in children are most likely due to benign
23

Vertigo

24
Table 2.2.

Dizziness or light-headedness as key symptoms

Cause

Associated signs and symptoms

Trigger

Mechanism I disease

Presyneopal dizziness
Orthostatic
hypotension

Blurred vision, muffled hearing,


pallor, weakness

- Standing up fram supine or sitting


position
- Antihypertensive or other drugs
- Diuresis, 1055 of water, or haemorrhage

Diminutions of intravascular volumes,


venous pooling, decline of blood
pressure with general cerebra I
ischaemia

Vasovagal attack

Blurred vision, muffled hearing,


pallor, weakness

- Strang emotional reactions


- Abrupt fear, pain, or vertigo

Parasympathetic hyperactivity in
limbic system and medullary
vasodepressor centre

Cardiac arrhythmia
(other more severe
causes: myocardial
infarcts, congestive
heart failure,
valvular disorders,
hypertensive crisis,
see also Table 2.3)

Palpitations, pounding heart

- Emotional or physical stress

Decline in cardiac output and blood


pressure with cerebral ischaemia

Emotional stress

Lowering of carbon dioxide with


cerebral vasoconstriction

Psyehosomatie dizziness
Hyperventilation
Sighing, anxiety, tachycardia,
paraesthesia of extremities and
periorally, lump in the throat,
tightness in the ehest, carpopedal
spasms
Panic attack (p. 458)

Fear of dying, losing contral, or going


Situationally bound or predisposed,
crazy,pounding heart, sweating,
with or without agoraphobia
trembling,nausea or abdominal distress

Anxiety disorder

Disturbance of concentration,
restlessness, sensation of hunger,
tremor, sweating, pallor,palpitations,
stupor

Inadequate food or insulin intake,


alcohol consumption

Diabetes mellitus,
insulin-secreting tumour

Cloudiness, drawsiness, 1055 of


concentration, ataxia, dysarthraphonia,
saccadic pursuit, gaze-evoked
nystagmus

Excess intake of drugs (barbiturates,


benzodiazepines, antiepileptics),
alcohol, or other substance abuse

Intoxications

Agoraphobia (see p. 459),


acraphobia (see p. 460),
phobie postural vertigo
(see p. 469)
Metabolie dizziness
Hypoglycaemia

(see also hyperventilation


with hypocapnia and
alkalosis)
Intoxication
Drugs or alcohol

paroxysmal vertigo of childhood (p. 376), a migraine


equivalent. In adults, short attacks of rotatory vertigo may occur in Meniere's disease (p. 83), basilar
migraine (p. 329), or transient vertebrobasilar
ischaemia (p. 307). A Meniere's attack usually lasts
for periods of up to a few hours with associated
hearing loss and tinnitus; Meniere's disease is rare,
but can occur in children. Episodic vertigo is only
rotatory when it involves semicircular canal function
(p.28); it may manifest as to-and-fro vertigo, if
otolithic function (p. 29) is involved, for example, in
perilymph fistulas (p. 100). Table 2.5 summarises

both frequent and rare disorders that may cause


recurrent episodic vertigo.

Sustained (rotatory) vertigo


Sustained (rotatory) vertigo occurs either with acute
unilateral peripheralloss of vestibular function or
with pontomedullary brainstem lesions near the
vestibular nudei. Vestibular neuritis is the most frequent cause, and its diagnostic hallmark is unilateral
hyporesponsiveness to thermal irrigation. Table 2.6

2S

Approaching the patient


Table 2.3. Classification of syncope

Table 2.4. Causes of orthostatic hypotension

Cardiac disease
Rhythm disturbances
Ventricular and supraventricular tachycardia
Bradycardia
Sinus node dysfunction
Second-degree AV block
Third-degree AV block
Pacemaker malfunction

Diminution of intravascular volume


Venous pooling (prolonged periods of physical inactivity, varicose veins)
Volume depletion (excessive diuresis - drugs, insufficiency of adrenal
glands, diabetes insipidus -Ioss of water due to enteropathy,
haemorrhage, excessive sweating)

Underlying structural heart disease


Acute myocardial infarction
Aortic stenosis
Hypertrophie obstructive cardiomyopathy
Mitral valve stenosis
Atrial myxoma
Pulmonary artery stenosis
Pulmonaryembolus
Pulmonary hypertension
Tetralogy of Fallot
Non-cardiac disease
Vascular disorders
Vasovagal syncope
Orthostatic hypotension
Carotid sinus syndrome
Situational faint
Micturition syncope
Defecation syncope
Cough syncope
Deglutition syncope
Postprandial hypotension
Cerebrovascular disturbances
Transient ischaemic attack
Subclavian steal syndrome
Takayasu's disease
Aortic dissection
Metabolie disturbances
Hypoglycaemia
Disturbances in electrolyte levels predisposing to arrhythmias
Hyperventilation

Drug-induced
Antihypertensive agents
Nitrates
Neuroleptics
Tricyclic antidepressants
Sedatives
Levodopa
Diuretics
Calcium channel blockers
Antiarrhythmics
Autonomie failure
Peripheral neuropathy (especially in diabetes mellitus)
Idiopathic autonomie neuropathy (Bradbury-Eggleston syndrome)
Shy-Drager syndrome (multisystem degeneration)
Parkinson's disease
Wernicke's encephalopathy
Spinal cord disease (tabes dorsalis, syringomyelia)
From Laicher and Linzer (1996).

Table 2.S.

Disorders that can cause recurrent episodic vertigo

Peripheral
Peripheral and/or central
labyrinth/eighth nerve

----_._---_.

__

._----------

Meniere's disease

Basilar migraine

Vestibular epilepsy

Vestibular paroxysmia

Benign paroxysmal
vertigo of childhood

Room-tilt illusion

Perilymph fistula

Benign recurrent vertigo

Paroxysmal
ataxia/dysarthria
(MS)

Benign paroxysmal
positioning vertigo

Vertebrobasilar ischaemia

Familial episodic
ataxia I, 11

Cogan's syndrome

Anterior-inferior
Paroxysmal ocular
cerebellar artery ischaemia tilt reaction

Unexplained syncope
From Laicher and Linzer (1996).

Central vestibular

Syphilitic labyrinthitis
Vestibular atelectasis
Otosclerosis
Hyperviscosity
syndrome
Acoustic neurinoma
Cerebellopontine
angle cyst
Vestibular neuritis

Vertigo

26
Table 2.6. Disorders that can cause sustained rotational and/or to-andfra vertigo due to unilateral peripheral vestibular dysfunction

Positional/positioning vertigo

Infections
Viral
Vestibular neuritis
Viral neurolabyrinthitis
Herpes zoster oticus
Human immunodeficiency virus
Spumaretrovirus
Bacterial
Tuberculous labyrinthitis
Syphilitic labyrinthitis
Chlamydiallabyrinthitis
Lyme borreliosis
Bacterial meningitis
Cholesteatoma

Positional/positioning vertigo is due to canalolithiasis


in the posterior semicircular canal (p. 251) in the
majority of the patients presenting with this condition. It is correctly called benign paroxysmal
positioning vertigo (BPPV), because it is the rapid
positioning manoeuvre rather than the change of
the position of the head relative to the gravity vector,
which causes the manifestation. The diagnostic clue
is the characteristic transient rotatory nystagmus
beating toward the undermost ear when the sitting
patient is rapidly tilted sideways or is brought into a
slightly rotated head-hanging position. Another
mechanism of positional vertigo may be neurovascular cross-compression (vestibular paroxysmia or
"disabling positional vertigo"; p. 117). All central
forms of positional vertigo involve the region
around the vestibular nuclei and a neuralloop to the
cerebellar vermis. The differential diagnosis of positional vertigo is depicted in Section D, Table 1
(p. 249). Sometimes patients experience vertigo only
during particular lateral head rotations (Table 2.7);
others report head motion intolerance with distressing unsteadiness and oscillopsia (Table 2.8).

Autoimmune inner ear disorders (see Table 9.3)


Tumour
Acoustic neurinoma
Meningeoma
Epidermoid eysts
Glomus body tumour
Metastatie carcinoma

Vascular
Labyrinth infaretion
Vertebrobasilar eetasia
Hyperviscosity syndrome

Traumatic
Temporal bone fraeture
Labyrinthine concussion
Post-traumatic otolith vertigo
Perilymph fistula

latrogenic
Post-ear surgery
Transtympanic gentamicin treatment

Table 2.7.
rotation

Vestibular paroxysmia (p. 117) (disabling positional vertigo)


Rotational vertebral artery oeclusion (p. 296)
Rotational compression of the eighth nerve by cerebellopontine angle
mass (p. 124)
Carotid sinus syndrome (Table 2.3)

Table 2.8.
anee

summarises other, less frequent, causes of unilateral


peripheralloss of vestibular function. Again, the perceived vertigo is not necessarily rotatory, especially
when otolithic function is affected as, for example, in
post-traumatic otolith vertigo (p. 349). Within a few
weeks, the initial vestibular tone imbalance (due to
the unilateralloss) is (re)equalised by central compensation (p. 56). Thus, chronic unilateral vestibular
lesions are mostly asymptomatic. Differential diagnosis of the pathologies of sustained central vertigo
involves all acute processes of the intra-axial
infratentorial structures (involving the root entry
zone of the eighth nerve or the vestibular nuclei)
such as multiple sclerosis, tumours, or brainstem
infarctions. Clues for the diagnostic steps are mostly
given by additional brainstem signs, in particular
ocular motor abnormalities that exceed those
expected of a peripherallabyrinthine loss.

Dizziness, vertigo, or disequilibrium elicited by lateral head

Vertigo and oscillopsia with particular head motion intoler-

Bilateral vestibulopathy
Oeular motor disorders (defeetive VOR)
Vestibular paroxysmia ("disabling positional vertigo")
Benign paroxysmal positioning vertigo
Central positional/positioning vertigo
Vestibuloeerebellar ataxia
Perilymph fistula
Post-traumatie otolith vertigo
Carotid sinus syndrome
Rotational vertebral artery oeclusion
Intoxieation (e.g. alcohol, phenytoin)

Oscillopsia
Patients with involuntary ocular oscillations
(acquired pendular nystagmus, downbeat and
upbeat nystagmus) not only report a decline of visual

27

Approaching the patient

acuity, but also apparent motion of the visual scene Table 2.10. Vertigo syndromes that eause eombinations of vestibular
(oscillopsia). Patients with extraocular muscle pare- and auditory dysfunetion (see also Table 2.11)
sis or defects of the vestibulo-ocular reflex are often Meniere's disease
unable to recognise faces or to read while walking; Perilymph fistula
they can also report oscillopsia. Either the deficiency Tumours of eerebellopontine angle and temporal bone
of compensatory eye movements (due to an inappro- Vestibular paroxysmia (neurovaseular eompression)
priate vestibulo-ocular reflex) or the deficiency of Ear trauma
visual fixation (due to ocular oscillation) causes Autoimmune inner ear disease (e.g. Cogan's syndrome)
undesired retinal image motion with disturbing Otosclerosis
Ear infaretion
oscillopsia and sometimes unsteadiness (p.430). Neurolabyri nthitis
Conditions that may cause oscillopsia with or with- Cholesteatoma
out head motion are depicted in Tables 2.8 and 2.9.
Congenital malformation

Vertigo associated with auditory dysfunction


The presence of dizziness, vertigo and disequilibrium
combined with sensorineural hearing loss or tinnitus narrows down the differential diagnosis to certa in peripheral vestibular disorders (summarised in
Table 2.10). The rare central vestibular disorders that
may manifest with audiovestibular symptoms are
vestibular epilepsy or caudal brainstem dis orders
such as in multiple sclerosis. Audiovestibular dysfunction associated with interstitial keratitis indicates infectious or auto immune disease (Table 2.11).
Congenital unilateral or bilateral vestibular disorders (Table 2.12) may be combined with sensorineural hearing loss.

Dizziness or to-and-fro vertigo and postural


imbalance
Dizziness, postural imbalance and unsteadiness of
gait are non-specific but frequently described symptoms. Differential diagnosis on the basis of such
symptoms is the most difficult, because central and
peripheral vestibular disorders but also non-vestibular
syndromes such as visual vertigo, presyncopal faintness, or psychosomatic dis orders are all possible
Table 2.9. Oscillopsia as a major eomplaint
Without head motion
Congenital nystagmus (dependent on
direetion of gaze)
Downbeat nystagmus
Upbeat nystagmus
Aequired pendular nystagmus
Periodie alternating nystagmus
Opsoclonus
Ocular flutter
Superior oblique myokymia
Paroxysmal oeular tilt reaetion
Spasmus nutans (infants)
Voluntary nystagmus
Spontaneous vestibular nystagmus

Various hereditary disorders with vestibulocoehlear involvement (p. 378)


Vestibular ateleetasis
Hyperviseosity syndrome
Vestibular epilepsy

Table 2.11. Differential diagnosis of audiovestibular dysfunetion plus


interstitial keratitis
Cogan's syndrome
Congenital/aequired syphilis
Chlamydial infeetions
Tubereulosis
Sarcoidosis
Viral infeetions (Herpes zoster, mumps, rubella, rubeola)
Vaseulitis (polyarteritis nodosa, temporal arteritis, Wegener's
granulomatosis)
Vogt-Koyanagi-Harada disease
Modified after Haynes et al. (1980).

Table 2.12. Congenital vestibular dysfunction


Familial episodic ataxia types land 11
Familial vestibular areflexia
Various hereditary disorders with vestibulocoehlear involvement
(p.378)
Congenitallabyrinth malformation (e.g. Mondini's dysplasia, perilymph
fistula)
Embryopathie labyrinth malformation (rubella, eytomegalovirus,
thalidomide)
Congenital syphilitie labyrinthitis
Congenital upbeat/downbeat nystagmus (e.g. Arnold-Chiari
malformation)
Craniocervical malformations

With head motion


See Table 2.8

diagnoses. Two conditions should be mentioned in


which the subjective complaint of severe postural
instability is at variance with the clinical finding of
apparently normal vestibular function: the otolithic
type of perilymph fistula (p. 100) and the posttraumatic otolith vertigo (p. 349). The most likely
differential diagnosis of the latter is phobic postural
vertigo (p. 469), which also manifests as a sequel to
initial vestibular dysfunction. The combination of
oscillopsia with head movements and unsteadiness
of gait is typical of bilateral vestibulopathy (p. 127).

28

Vertigo

In order to differentiate peripheral and central Table 2.14. Dizziness and unsteadiness due to intoxication: typical
vestibular causes among this group, the direction of clues, signs and symptoms
fall is significant (see Vestibular falls, p. 13; Fig. 1.6, Fluctuations of
Table 1.3, Table 1.4, Table 2.13). Fluctuations of the - dizziness, drowsiness, confusion, disorientation
syndrome associated with ocular motor abnormali- - memory and cognitive deficits
ties and cerebellar ataxia, unusual sleeping patterns, - emotionallability
and repeated unexplained falls suggest intoxication - unsteadiness and gait ataxia
as the cause (Table 2.14). Experience has taught us - dysarthrophonia
when to suspect a psychiatrie cause of the condition - saccadic pursuit eye movements and gaze-evoked nystagmus
rather than otoneurological or neuro-ophthalmolog- - positional nystagmus
- miosis and mydriasis and other autonomic signs
ical causes (Table 2.15). Typical signs and symptoms
are based mainly on the dissociation of the severity Unusual sleeping patterns
of subjective complaints and the normal findings on - sleeping at various times during the day
- prolonged daily sleep phases (> 8 h)
clinical examination, the situational dependence of
dizziness and unsteadiness with inadequate avoid- Repeated unexplained falls in the ho me environment
an ce behaviour, and transient improvement after Discrepancy between the complaints of the relatives and the obvious
dissimulation of the afflicted patient
alcohol intake.
Additional tables for differential diagnosis of central vestibular dis orders (Chaps. 10-15), traumatic
vertigo (Chaps. 22 and 23), vertigo in childhood
(Chap. 26), vertigo in the elderly (Chap. 27), drugs Table 2.1 S. Dizziness, vertigo and disequilibrium as a psychosomatic
and vertigo (Chap. 28), visual vertigo (Chap. 29), manifestation: typical signs and symptoms
psychogenic vertigo (Chaps. 31 and 32) appear in the Dissociation of
individual chapters.
- subjective imbalance and objective balance skills
Table 2.13. Dizziness and postural imbalance
Vestibular
Predominant
fore-aft instability

Predominant
lateral instability

Multidirectional
instability

Downbeat nystagmus
Upbeat nystagmus
Alcoholic spinocerebellar
(anterior lobe) degeneration
Bilateral vestibular failure
Lateropulsion in Wallenberg's
syndrome
Ocular tilt reaction
Thalamic astasia
Corticallateropulsion ("pusher")
Otolithic, vestibular nuclei
Vestibulocerebellar dysfunction

Visual

Visual field defects


Visual acuity and refraction
anomalies
Ocular motor disorders

Somatosensory

Polyneuropathy
Cervical vertigo
Dorsal spinal cord lesions

Cerebellar

Cerebellar degeneration
Cerebellar lesions
Intoxication

Cardiovascular

Presyncopal

Psychogenic

Phobic postural vertigo


Hyperventilation syndrome
Panic attack
Agoraphobia
Acrophobia

- fear of losing control of stance, gait, or driving ability and missing


history of the like
- moderate, frequently vague visual-vestibular symptoms and excess
anxiety
- complaints of rotatory vertigo and missing associated nystagmus or
ocular motor abnormalities
Situation-specific occurrence of major symptoms
Inadequate avoidance behaviour with rapid conditioning and
generalisation
Transient improvement of symptoms after drinking small amounts of
alcohol or during periods of distraction bya particular physical activity
(Obsessive-compulsive type personality, labile affect, mild depression,
obvious suffering)

Semicircular canal vertigo and mixed


canal-otolith vertigo
Most vestibular syndromes involve semicircular
canal and otolithic function for several reasons:

The different receptors for perception of angular


and linear accelerations are housed in a common
labyrinth.
Their peripheral (eighth nerve) and central (e.g.
medial longitudinal fascicle) pathways take the
same course.
There is a convergence of otolith and semicircular canal input at all central vestibular levels,
from the vestibular nuclei to the vestibular
cortex.

29

Approaching the patient

Thus, most vestibular syndromes are mixed as


regards otolithic and canal function. A peripheral
prototype of such a mixture is vestibular neuritis
caused by inflammation of the superior division of
the vestibular nerve that subserves the horizontal
and the anterior semicircular canals and the maculae of the utricle and the anterosuperior part of the
saccule (p. 73). A central prototype is Wallenberg's
syndrome which involves the medial and superior
vestibular nuclei where otolith and canal input converge. This typically causes ocular and body lateropulsion and torsional spontaneous nystagmus
(p. 309).
It is, however, possible to selectively stimulate single canals by caloric irrigation of the external
acoustic canal. The prototype of a semicircular canal
disease is benign paroxysmal positioning vertigo of
the posterior or horizontal canal (p. 251). Typical
signs and symptoms of semicircular canal vertigo
are

rotational vertigo and deviation of perceived


straight-ahead,
spontaneous vestibular nystagmus with oscillopsia,
postural imbalance with Romberg fall and pastpointing, and
nausea and vomiting, if severe.

The three-dimensional (3D) spatial direction of nystagmus and vertigo depends on the spatial plane of
the affected semicircular canal and on whether the
dysfunction is caused by ampullofugal or ampullopetal stimulation or a unilateral loss of afferent
information. Malfunction of a singular or more than
one semicircular canal can be detected by threedimensional analysis of spontaneous nystagmus
(Straumann and Zee 1995; Bhmer et al. 1997; Fetter
et al. 1997) or perception of rotation (von Brevern et
al. 1997). Central vestibular syndromes may take
precedence over semicircular canal or otolith type.
They are best classified according to the three major
planes of action of the vestibular ocular reflex: yaw,
roll and pitch (p. 169). To put it simply, "dynamic",
rotatory vertigo and nystagmus represent (angular)
canal function, whereas "static" ocular tilt reaction,
body lateropulsion, or tilts of perceived vertical represent (linear) otolith function.
Galvanic stimulation affects the entire eighth
nerve with the semicircular canals and otoliths (Fig.
2.1; Zink et al. 1997, 1998). Functional MRI during
galvanic stimulation shows three different sensory

systems that are activated in the insula-thalamic


region: the vestibular, auditory and nociceptive (Fig.
2.1; Bucher et al. 1998).

Otolithic vertigo
Although the pathophysiology of otolithic dysfunction is poorly understood, a disorder of otolithic
function at a peripheral or centrallevel should be
suspected if a patient describes symptoms of falls,
sensations of linear motion, or tilt, or else shows
signs of specific derangements of ocular motor and
postural orienting and balancing responses (Gresty
et al. 1992). A significant number of patients presenting to neurologists have signs and symptoms that
suggest disorders of otolithic function. Nevertheless,
diseases of the otoliths are poorly represented in our
diagnostic repertoire (Table 2.16). Of these, posttraumatic otolith vertigo (p. 349; Brandt and Daroff
1980) may be the most significant; the rare otolith
Tullio phenomenon may be the best studied (p. 107;
Dieterich et al. 1989; Fries et al. 1993). Other examples
are vestibular drop attacks (Tumarkin's otolithic
crisis) and a number of central vestibular syndromes
that indicate tone imbalance of graviceptive circuits
(skew deviation, ocular tilt reaction, lateropulsion,
room-tilt illusion), some of which manifest without
the sensation of dizziness or vertigo.

Paroxysmal vertigo
Vertigo and other vestibular syndromes may result
from pathological excitation of various vestibular
structures: the labyrinth, the vestibular nerve,
the vestibular nuclei and their ascending pathways to the thalamus and the cortex (Table
2.17; Brandt and Dieterich 1994a). Three features
are typical for most paroxysmal vestibular
syndromes:
1. short duration of paroxysms (seconds to minutes)
2. frequent repetitive occurrence that is spontaneous or triggered by various stimuli
3. the direction of vertigo, nystagmus and falls is
opposite to that of lesional dysfunction of the
affected vestibular structure.

30

Vertigo

Galvanic Vestibular Stimulation


1.0 - 5.0 mA
5s

10s

Mastoid

G4/G5

DDDO

15s

D
BB

00000=0
Gurrent Source

ccw 2

[1 2 mA 1

cw

QO[

T ccw

::~

3 mA

4 mA

cw 0
ccw 4
30

b
Fig.2.1.

20s

Approaching the patient

31

Fig.2.1. Galvanic stimulation of the entire eighth nerve.


a Stimulation at the mastoid and at the (4/(5 level for comparison of non-vestibular effects. b Two patterns of eye movements during
galvanic vestibular stimulation with different current intensities. The two upper traces show torsional eye movements during stimulation with 2 mA and 3 mA DC current, respectively. The three lower traces show torsional, vertical, and horizontal eye position during
stimulation with 4 mA DC current. On the left, a typical response pattern as seen in six of seven subjects with a static torsional deviation (otolith stimulation) accompanied by one or two nystagmus beats with current intensities of 3 mA and 4 mA. On the right, eye
movement patterns as seen in one of seven subjects with horizontal-torsional nystagmus (semicircular canal stimulation) with all current intensities applied. Slight horizontal nystagmus can be seen on the lowest trace on the right. The absence of a vertical deviation
and nystagmus can be explained by the counter-directed vertical components of the anterior and posterior semicircular canals. Thus,
galvanic stimulation affects the entire eighth nerve. (Fram Zink et al. 1998.) c fMRI during galvanic stimulation. Magnified cortical and
subcortical activation maps superimposed on the corresponding coronal T2* -weighted anatomical images of three sections of a 31year-old subject (TR/TE=63/30 ms, a=l 0). The superimposed activation maps are associated with galvanic stimulation at the mastoid
level (a, c, e) and galvanic stimulation at the (4/(5 level (b, d, f) for contral. The colour-coded correlation coefficient scale ranges fram
0.5 (red) to a maximum of 1.0 (yellow). Stimulation at the mastoid level (vestibular stimulation and cutaneous pain stimulation) caused
activation in the medial part of the insula (MI; a, c and e). the posterior part of the insula (PI; a, c and e), the transverse temporal gyrus
(TTG; a and cl. the anterior median thalamus (AT; c and e) and the posterior median thalamus (PT; c and e).ln contrast, stimulation at
the (4/(5 level (cutaneous pain stimulation) was associated with activity in the medial part of the insula (b, d and f) and in the anterior
median thalamus (d and f) only. Thus, galvanic stimulation at mastoid level activates cortical areas of three different sensory systems
in the insula-thalamic region: the vestibular, auditory and nociceptive systems. (From Bucher et al. 1998.)

Vertigo

32
Table 2.16. Peripheral and central vestibular syndromes affecting otolith function

Disorder

Signs/symptoms

Mechanism

Head motion intolerance, gait ataxia, to-and-fro


vertigo, tilt of perceived vertical, skew deviation, lateropulsion

Dislodged otoconia cause unequal heavy


loads with "graviceptive" tone imbalance

Vestibular drop attacks (Tumarkin's


otolithic crisis in Meniere's
disease)

Sudden falls, sensation of being pushed to the ground,


Meniere's triad

Sudden changes in endolymphatic fluid


pressure with inappropriate otolith
stimulation causing reflex-like
vestibulospinalloss of postural tone

Endolymphatic hydrops

Episodic to-and-fro vertigo, unsteadiness, Meniere's disease

"Floating labyrinth~ deformation or


pressure changes in the membranous
labyrinth

Perilymph fistula
(otolith type)

To-and-fro vertigo, gait


ataxia with sneezing, coughing, or
physical exercise, positive fistula signs (e.g. Valsalva's
manoeuvre)

Perilymph leakage,
abnormal elasticity of the bony labyrinth with
irritative otolith stimulation during head
motion, intracranial pressure changes

Vestibular atelectasis

Episodic to-and-fro vertigo, gait ataxia

Collapse of the walls of the ampulla and


utricle

Otolith Tullio phenomenon

Sound or pressure-induced paroxysms of perceived tilt,


oscillopsia, skew deviation and lateropulsion

Inadequate mechanical stimulation of otolith


by hypermobile stapes footplate (stapedius
reflex) caused by loud sounds

Triad of head tilt, skew deviation, and ocular torsion


associated with perceived tilt

"Graviceptive" tone imbalance due to


unilateralloss or irritation of otolithic
function

Paroxysms of vertical and torsional diplopia, perceived


tilt, head and body lateropulsion

Neurovascular cross-compression of the


(utricular?) nerve with ephaptic spreading

Paroxysmally perceived tilts and body falls with or


without ocular motor abnormalities

Epileptic discharges in vestibular cortex

Corticallateropulsion

Body tilt and tilts of perceived vertical

Cortical "graviceptive" tone imbalance with


acute lesions of the parieto-insular vestibular
cortex

Room-tilt illusion

Transient illusions of upside-down vision or apparent 90


tilts of the visual scene

Cortical mismatch of visual and otolithic 3D


maps of spatial orientation

Lateropulsion and tilt of perceived vertical

"Graviceptive" tone imbalance with acute


lesions of vestibular subnuciei

See above

See above

Modulation of nystagmus by changes in head position

Modulation of vestibulo-ocular reflex tone in


pitch by changes in graviceptive input

Peripheral vestibular

Labyrinth
Post-traumatic
otolith vertigo

Eighth nerve/or labyrinth


Ocular tilt reaction

Eighth nerve
Vestibular (otolithic)
pa roxysmia
Central vestibular

Cortex
Vestibular epilepsy

Thalamus
Thalamic astasia
Brainstem
Ocular tilt reaction
Lateropulsion
Room-tilt illusion
Upbeat/downbeat
nystagmus provoked or
modulated by changes in
head position

33

Approaching the patient

Table 2.17. Synopsis of paroxysmal vertigo and other vestibular syndromes resulting from pathological excitation rather than 1055 of function induced by a
lesion of vestibular structures
Syndrome
~

~~---

Vestibular site

Cortex
Vestibular epilepsy

Mechanism
~----

-~

Facultative trigger

Reference

~----

Vestibular cortex

Simple or complex
partial (vestibular)
sensorimotor seizures

Epileptic triggers
Electrical stimulation

Foerster 1936
Penfield and Jasper 1954
Schneider et al. 1968

Epileptic nystagmus

Vestibular or visual or
temporo-occipital cortex

Simple or complex
partial sensory (vestibular,
visual or optokinetic ?)
seizures

Visual input
Eccentric gaze
Electrical stimulation

Stodieck et al. 1990


lusa etal. 1990
Stolz et al. 1991

Room-tilt illusion

Parieto-occipital
or frontal cortex

Mismatch of cortical
matching of visual
and vestibular 3D coordinate maps

Vestibular tone imbalance

Ropper 1983
liliket et al. 1996
Brandt 1997

Pontine tegmentum,
brachium conjunctivum

Ephaptic (non-synaptic)
spreading activation of
demyelinated adjacent axons

Rising or locomotion
Hyperventilation

Andermann et al. 1959


Espir and Millac 1970
Osterman and Westerberg
1975

Rostral midbrain,

Ephaptic (non-synaptic)
spreading activation or
stimulation of unilateral
graviceptive pathways of
VOR in roll plane

Brainstem
Paroxysmal dysarthria,
vertigo, and ataxia
in multiple sclerosis
Paroxysmal ocular tilt
reaction

Paroxysmal vertigo/
nystagmus with lateral gaze
Familial episodic
ataxia land 11

Interstitial nucleus of Cajal

Electrical stimulation

Rabinovitch et al. 1977


Hedges and Hoyt 1982
Lueck et al. 1991

Prolonged lateral gaze

Bttner et al. 1987

Autosomal dominant
potassium channel disease
EA 1, calcium channel
disease EA2

Exertion, fatigue,
emotional stress, alcohol

Brunt and van Weerden 1990


Griggs and Nutt 1995
Brandt and Strupp 1997

Lateral medulla,
vestibular nuclei ?
Cerebellum?
brainstem?

Convergence -evoked
nystagmus

Various ocular motor and


vestibular brainstem
pathways

Ephaptic spreading (?L


unmasks or accentuates
ocular motor or vestibular
tone imbalance

Attempted convergence

Sharpe et al. 1975


Cox et al. 1981
Oliva and Rosenberg 1990

Room-tilt illusion

Vestibular nuclei,
caudal brainstem

Mismatch of cortical matching


of visual and vestibular
3D coordinate maps

Vestibular tone imbalance

Ropper 1983
liliket et al. 1996
Brandt 1997

Eighth nerve root

Neurovascular cross-compression with demyelination


and ephaptic spreading

Individual head positions

M011er et al. 1986


Jannetta et al. 1984
Brandt and Dieterich 1994b

Eighth nerve root

lumour compression
(petrous bone, cerebellar
pontine angle) of the eighth
nerve

Hyperventilation

Leigh and Zee 1991

Semicircular canals
or otoliths

Inadequate sound-induced
mechanical stimulation of
otoliths or canals in
perilymph fistulas or luxation of the stapes footplate

Loud sounds
(stapedius reflex),
Valsalva manoeuvre

lullio 1929
Deecke et al. 1981
Dieterich et al. 1989

Posterior>
horizontal> anterior
semicircular canal

Canalolithiasis with a
free-floating "heavy"
clot within the endolymph
of the canal

Head positioning relative


to gravitational force

Brandt and Steddin 1993


Baloh et al. 1993

Semicircular canals

Buoyancy mechanism based


Head position relative
on whether the specific
to gravitational force
gravity of the cupula is greater
or less than that of the endoIymph

Vestibular nerve
Vestibular paroxysmia
(disabling positional vertigo)
Hyperventilation-evoked
paroxysmal nystagmus/
vertigo

Labyrinth
lullio phenomenon

Benign paroxysmal
positioning vertigo (BPPV)

Positional alcohol vertigo/


nystagmus (PAN)

Modified from Brandt and Dieterich (1994a).

Barany 1911
Aschan et al. 1956
Money et al. 1974

34

Vertigo

Neuro-ophthalmological and
otoneurological evaluation

Fig.2.2. Clinical examination with Frenzel's glasses. The magnifying lenses (+ 16 dioptres) have light inside to prevent visual fixation, which could suppress spontaneous nystagmus. Frenzel's
glasses enable the clinician to better observe spontaneous eye
movements. Examination should include spontaneous and gazeevoked nystagmus, head-shaking nystagmus (instruct the
patient to rotate his head about 20 times and observe eye movements following head shakingl, positioning (see p. 253) and
positional nystagmus (see p. 292)' as weil as hyperventilationinduced nystagmus. Spontaneous nystagmus indicates a tone
imbalance of the vestibulo-ocular reflex which may be central or
peripheral; when peripheral- as in vestibular neuritis - it is typically damped by visual fixation. Head-shaking nystagmus is caused
by an asymmetry of velocity storage which occurs after both
peripheral and central vestibular lesions.

Fig. 2.3. Clinical examination with Frenzel's glasses and a


Politzer balloon. Changes in middle-ear pressure by applying positive or negative pressure to the tympanic membrane with a
Politzer balloon; noise or tragal compression may induce nystagmus in patients with perilymph fistula. Nystagmus mayaiso be
observed during Valsalva's manoeuvre, swallowing, or coughing.

Fig.2.4. Ophthalmoscopy (if the other eye is covered, visual fixation is prevented) is a sensitive method for detecting spontaneous nystagmus (Zee 1978) even with low, slow phase
velocities/frequencies or square-wave jerks (small saccades
[OS-5 0 ] that displace the eye from the primary position; it occurs,
e.g. in progressive supranuclear palsy or certain cerebellar syndromes) while checking for movements of the optic disc. Since
the retina is behind the centre of rotation of the eyeball, the
direction of any observed vertical or horizontal movement is
opposite in direction to that of the nystagmus, i.e. a down beat
nystagmus causes fast, upwardly directed movements of the
optic disc.

Approaching the patient

Fig.2.5. Measurement of head tilt. Look for abnormal head


postures. Tilt is observed especially in patients with either paresis
of the oblique eye muse/es, (e.g. in superior oblique palsy, the head
is turned to the non-affected side to lessen diplopia) or an ocular
tilt reaction due to a tone imbalance of the VOR in roll. Acute
lower medullary lesions (e.g. in Wallenberg's syndrome) and
peripheral vestibular lesions cause an ipsiversive head tilt, mesencephalic lesions, a contraversive head tilt. The head is usually
tilted toward the side of the lower eye.

35

Fig.2.6. Cover tests: detection of misalignments of the visual


axes. The unilateral cover test reveals heterotropia, i.e. a misalignment of the visual axes when both eyes look at a single target.
First the patient has to fixate either a near target (at a distance of
30 to 40 cm) or one 5 to 6 m distant. Subsequently one covers
the right eye and looks for movements of the uncovered left eye.
If a movement of the left eye is detected, the patient has a leftsided heterotropia.lf the left eye moves from the inside outward,
there is an esotropia; if it moves from the outside inward, there is
an exotropia; if it moves from above downward, there is a hypertropia; if it moves from below upward, there is a hypotropia. The
cover is then removed and the left eye is covered. The alternating
cover test reveals the maximal deviation of tropia or phoria . lt is
also useful for detecting skew deviation (part of the ocular tilt
reaction), a vertical misalignment of the eyes that cannot be
explained by an ocular muscle/ nerve palsy. One looks for vertical
corrective movements, when the cover is switched from one eye
to the other. ln contrast to fourth cranial nerve palsy, skew deviation changes little with different directions of gaze. Latent congenital nystagmus is a jerk nystagmus that is absent when both
eyes fixate; it appears when one eye is covered and often
changes direction depending on which eye is uncovered.

36

Vertigo

Fig.2.7. Clinical examination of the eyes in nine different positions to evaluate ocular alignment, fixation deficits, nystagmus, range
of movement and gaze-holding abilities. The examination can be performed using an object (Ieft) or an examination lamp (see Fig.
2.8).ln primary position look for (1) abnormal eye movements such as nystagmus (e.g. peripheral vestibular: horizontal-rotatory, suppressed by fixation; central vestibular: vertical (upbeat, down beat), horizontal or torsional, poorly suppressed or even increasing with
fixation; congenital: usually horizontal, variable in frequency and amplitude, increasing with fixation); square-wavejerks (small saccades [0.5-5] that cause the eyes to move from the primary position, e.g. in progressive supranuclear palsy or certain cerebellar syndromes); ocular flutter (intermittent bursts of horizontal oscillations); or opsoc/onus (combined horizontal, vertical and rotatory
oscillations); the latter two may have different aetiologies, for example, encephalitis, tumours, or drugs/toxins and (2) misalignment of
the visual axes (see cover test, Fig. 2.6). Then establish the range of motion with ductions (one eye viewing) and versions (with both
eyes viewing) in the eight end-positions; this can indicate, e.g. ocular muscle or nerve palsy. Gaze-holding deficits can be evaluated in
eccentric gaze position (see Fig. 2.8) ..

Approaching the patient

Fig.2.8. Clinical examination of the eye positions/movements


using an examination lamp. The advantage of the lamp as
opposed to an object is that the reflected light on the eye can be
observed and thus ocular misalignments can be easily detected.
In addition, the patient can fixate with one or both eyes in the
end-positions. Gaze-evoked nystagmus should be observed when
the patient is fixating with both eyes.lt is most often a side effect
of medication/toxins such as anticonvulsants, hypnotics, or alcohol. Horizontal gaze-evoked nystagmus may be due to structural
lesions of the brainstem (vestibular nucleus and nucleus prepositus hypoglossi, i.e. the neural integrator), the flocculus, or the
medial vestibular nucleus. Vertical gaze-evoked nystagmus is
observed in midbrain lesions involving the interstitial nucleus of
Cajal. A dissociated horizontal gaze-evoked nystagmus (greater in
the abducting than the adducting eye) in combination with an
adduction deficit are the signs of internuclear ophthalmoplegia
due to alesion of the medial longitudinal fasciculus (MLF).
Downbeat nystagmus increases usually in eccentric gaze position.
Ifthe patient returns the eyes to the primary position after prolonged maintenance of eccentric gaze, a transient nystagmus
may ~ppear. with slow phases in the direction of the previous eye
position. Thls so-ca lied rebound nystagmus usually indicates ce rebellar lesions.

37

Fig. 2.9. Left: Clinical examination of smooth pursuit: the


patient is asked to track visually an object moving slowly in horizontal and vertical directions (10 to 20'/s) with the head stationary.. Look for corrective (catch-up or back-up) saccades; they
Indlcate an Inappropriate smooth pursuit gain. Many anatomical
structures (visual cortex, MT, MST, frontal eye fields, dorsolateral
pontine nuclei, cerebellum, vestibular and ocular motor nuclei)
~re involved in smooth pursuit eye movements, which keep the
Image of a moving object on the fovea. Therefore, impaired
smooth pursuit (reduced gain) is an unspecific finding, wh ich
may be further influenced by alertness, a variety of drugs, and
age. Moreover, vertical smooth pursuit is worse than horizontal
and ?ownward tracking, worse than upward. Marked asym~
metrtes of smooth pursuit, however, indicate a centrallesion;
strongly impaired smooth pursuit is observed in intoxications
and degenerative disorders involving the cerebellum or
extrapyramidal system. Reversed smooth pursuit is often found in
congenital nystagmus. Right: Vergence test ond near triad. Move a
target along the patient's midsagiUal plane from a distance of
about 50 cm toward the bridge of the nose. This causes vergence, accomodation and pupillary constriction, i.e. the near
triad. Neurons important for vergence have been found in the
mesencephalic reticular formation and the oculomotor nucleus.
This explains why disturbances of vergence occur in rostral midbrain lesions and tumours of the pineal region and thalamus, and
ar~ often associated with abnormalities ofvertical gaze. In certaln neurodegenerative disorders such as progressive supranuclear palsy, vergence mayaiso be impaired. Inborn defects of
accomodative-convergence synkinesis also accompany some
forms of childhood strabismus (concomitant strabismus).
Con.vergence-retraction nystagmus can be induced by having the
patient look at a moving optokinetic drum while its stripes are
gOlng downward or by having him make upward saccades.
Instead of quick upward phases, the patient makes rapid converge~t movem~nts associated with retractions of the eyeball.
Anlma.1 experiments have shown that lesions of the posterior
commissure cause convergence-retraction nystagmus. Spasm of
the near reflex is a voluntary convergence accompanied by pupillary constriction, which is a functional disorder that can mimic
bilateral abducens palsy.

38

Fig.2.10. Clinical examination of saccades. First observe spontaneous saccades to visual or auditory targets. Then ask the
patient to glance back and forth between two horizontal and
two vertical targets, keeping the head stationary. The velocity,
accuracy, conjugacy, and the initiation time of the saccade
should be observed. Normal individuals can immediately reach
the target with a fast single movement or one small corrective
saccade. Slowing of saccades - often accompanied by hypometric
saccades - is sometimes not caused bya structurallesion but is a
side effect of many types of medications/toxins and is also found
in neurodegenerative disorders.Slowing of horizontal saccades is
observed in brainstem lesions, e.g. of the ipsilateral paramedian
pontine reticular formation (PPRF); slowing of vertical saccades
may be due to a midbrain lesion affecting the rostral interstitial
MLF (riMLF) and is often observed in progressive supranuclear
palsy. Lesions of the cerebellum (especially the vermis) or cerebellar pathways may cause hypermetric saccades, followed by
corrective saccades that can be easily observed. For example, in
Wallenberg's syndrome, a saccadic overshoot toward the side of
the lesion is due to an interruption of the inferior cerebellar
peduncle; interruption of the superior cerebellar peduncle leads
to contra lateral hypermetric saccades.ln internuclear ophthalmoplegia (I NO) the adducting saccade is slower than the abducting saccade. Oelayed onset saccades are most often caused by
cerebral corticallesions.

Vertigo

Fig.2.11. The small optokinetic drum (or tape) allows combined (global) testing of optokinetic reflexive, smooth pursuit
movements and saccades in horizontal and vertical directions.lt
is especially helpful with uncooperative or drowsy patients. One
should look for asymmetries (e.g. between right and left in cerebral hemisphericallesions, smaller vertical than horizontal optokinetic nystagmus in supranuclear palsy), dissociarion of the two
eyes (diminished adduction in internuclear ophthalmoplegial,
and reversed pursuit in some patients with congenital nystagmus.
The optokinetic reflex is also helpful for disclosing psychogenic
blindness.

39

Approaching the patient

lesion of the
RIGHT labyrinth

healthy control

@
_

(1)

<;>

...
(2)

1.

HEAD ROTATION

CJj))

HEAD ROTATION

EYE MOVEMENT

EYE MOVEMENT

6{!J;;

(3)
~

(3)

I~

__________________________________

SACCADE

(2)/(3)
. :. . .(1-,-)~7
time
~

Fig.2.12. Clinical bedside testing of the horizontal vestibulo-ocular reflex by the Halmagyi-Curthoys test (Halmagyi and Curthoys
1988). Fast rotations of the head toward the side of the lesion reveal a dynamic deficit of the horizontal vestibulo-ocular reflex. In contrast to the healthy control a, the patient is not able to generate a fast contraversive eye movement and has to perform a corrective
(catch-up) saccade to fixate the target b. c Illustrates how to examine the patient.lt is important to instruct the patient to look carefully
at the examiner's nose and to apply brief, highly accelerated head turns to detect a unilateral peripheral vestibular deficit, e.g. due to
vestibular neuritis or acoustic neuroma.ln patients with chronic complete bilateral vestibulopathy, this test may surprisingly seem to
be normal (the cervico-ocular reflex partially substitutes for missing vestibular information); in these patients catch-up saccades may
be better detected by low frequency head oscillations. Cremer and colleagues (1998) have demonstrated that brief, unpredictable,
passive head impulses in a diagonal plane (midway between the frontal and sagittal planes) detect absent function of individual vertical semicircular canals.
(Part c on next page)

40

Vertigo

Fig.2.12c

Approaching the patient

41

Fig.2.13. Clinical testing of the visual fixation suppression of the vestibulo-ocular reflex (VOR). The patient is asked to fixate a target in
front of her eyes which moves horizontally and vertically with the patient's head. (Before the test the examiner must be sure that the
VOR is intact, for if there is no VOR, there is little to suppress).The examiner has to look for corrective saccades. Disturbed visual fixation
suppression of the VOR - almost always correlated with smooth pursuit abnormalities, as these two functions are mediated by common neural pathways - is often observed in lesions of the cerebellum (mainly flocculus or paraflocculus) or of the cerebellar pathways
and in progressive supranuclear palsy.lt mayaiso be caused by drugs (anticonvulsants and sedatives).

Fig. 2.14. Clinical examination of static balance.


Variations of the Romberg and one-Ieg stance: feet
together with eyes open or closed to eliminate visual
cues (upper left); on one foot at a time with the head in
normal position (10 wer left); with the head extended to
increase imbalance (upper right); and feet together with
eyes closed while the examiner distracts the patient by
writing numbers on her arm if a psychogenic disorder is
suspected (10 wer right). A further variation is the (sharpened) Romberg in tandem (not shown).One has to look
for excessive fore-aft, right-Ieft, or diagonal sway, e.g. a
peripheral-vestibular lesion causes ipsiversive falls,
up- and down beat nystagmus syndromes cause backwa rd fa IIs with the eyes closed.

42

Vertigo

Fig.2.15. Finger-pointing test. The patient is instructed to follow the finger of the examiner by rapidly pointing toward each
new position it takes. This test is more sensitive than the fingerto-nose test for ataxia, especially for dysmetria and hypermetria.

Fig.2.16. The Dix-Hallpike manoeuvre is used to diagnose benign paroxysmal positioning vertigo (BPPV) of the vertical semicircular
canals. The head of the patient is turned 45 to one side, and the patient is rapidly moved from a sitting to a supine position with the
head hanging over the end of the examination couch. In ca se of a BPPV of the left posterior semicircular canal, this manoeuvre will
induce a crescendo-decrescendo-like nystagmus with the upper pole of the eye beating toward the left ear (clockwise). This starts
with a latency of a few seconds, lasts less than 30 s, and reverses direction when the patient returns to a sitting position.

Approaching the patient

Fig. 2.17. Electronystagmography (ENG): typical electrode placement for monocular recording of horizontal and vertical eye
movements. The electrophysiological basis for the ENG is the
corneoretinal dipole that arises from the corneoretinal potential.
This potential has a magnitude of about 1 IlV and is oriented in
the direction of the long axis of the eye, with the retina being
negative and the cornea, positive. The difference in potential
between the active electrodes (in this case two horizontal electrodes for each eye, two vertical for the right eye) is DC-amplified.
The ENG allows non-invasive horizontal recordings of ca. 40
with an accuracy of ca. 1 and vertical recordings of ca . 20.
Major disadvantages are susceptibility to eyeblink artifacts, electromyographic activity, and unstable baseline; torsional eye
movements cannot be recorded with the ENG.

43

Fig.2.18. Electronystagmography: rotatory chair and rotatory


drum (with vertical stripes) with an apparatus that projects a
laser spot (above the patient). This arrangement allows recordings of eye movements under static conditions (e.g. test for
spontaneous or gaze-evoked nystagmus, saccades, pursuit, and
optokinetic nystagmus) and under dynamic conditions (perand postrotatory nystagmus, fixation suppression of the
vestibulo-ocular reflex), as weil as positional and positioning
testing and caloric irrigation (see Fig. 2.19)

44

Fig.2.19. Electronystagmography:caloric test. Bithermal caloric


irrigation is performed with 30C warm and 44C hot water, i.e.
7C below and above body temperature, respectively. The head
of the patient is tilted 30 upward, so that the horizontal semicircular canals are in the vertical plane, thus allowing optimal
caloric stimulation. The caloric test is the most widely used clinical test of the (horizontal) vestibulo-ocular reflex, because the
stimulus can be easily administered, each horizontal canal can be
stimulated individually, and it is weil tolerated by patients. The
maximum slow phase velocity (MSPV) is measured during each
irrigation. To compare the responsiveness of the right with the
left labyrinth, the "vestibular paresis formula" of Jonkees et al.
(1962) is most often used: (((R 30 oe + R 44 Oe) - (L 30 oe + L 44 Oe))!
(R 30 oe + R 44 oe + L 30 oe + L 44 Oe)) x 700, where, for instance, R
30C is the MSPV during caloric irrigation with 30C warm water.
Vestibular paresis is most often defined as > 25% asymmetry
between the right-sided and the left-sided responses (Honrubia
1994). Unilateral caloric hypo- or non-excitability is found most
often in peripheral vestibular lesions, e.g. in vestibular neuritis,
Cogan's syndrome, acoustic neuroma, or traumatic lesions of the
labyrinth or vestibular nerve. Central vestibular disorders, however, mayaiso cause caloric hypoexcitability (and mimic vestibular neuritis), especially lesions at the root entry zone of the
vestibular nerve ("fascicular lesions") due to plaques in multiple
sclerosis or lacunar infarctions.

Vertigo

Fig. 2.20. Posturography allows quantitative measurements


and documentation of postural stability on a force-measuring
platform, e.g. the so-ca lied sway path values (SP, mimin), the sway
direction, or a frequency analysis (Fourier power spectra) of the
sway. The SP is the length of the path described by the centre of
foot pressure during a given time (20 s), which is generated by
the inherent instability of a subject standing on a recording platform. It is approximated by the sum of the distances between
two consecutive sampling points in (a) the anteroposterior (sagittal = x) plane, i.e. sagittal sway (calculated as LI D.XIl, (b) the mediolateral (frontal = y) plane, i.e. frontal sway (calculated as (LI D.yl),
and (c) for both dimensions as the total SP (calculated as
( n/q D.X21 +L ID.y 2!l). Postural stability mayaiso be measured with
the patient standing on a foam-rubber padded platform (to
reduce ankle proprioception and, thereby, somatosensory input)
and with eyes closed (to eliminate visual input). These two conditions increase the particular sensorial weight of the vestibular
input in the multisensory control of postural balance. Some typical patterns of postural instability have been found: the cerebellar
pattern may be differentiated from that of spinal ataxia or postural
instability in basal ganglia disease; e.g.lesions of the upper vermial part ofthe anterior lobe - mainly observed in alcoholics lead to a typical 3 Hz anteroposterior body sway.

Approaching the patient

45

Fig. 2.21. Determination of subjective visual vertical (SW). The adjustment of the SW is measured in an upright position. The patient
looks into a hemispheric dome 60 cm in diameter. The surface of the dome extends to the limits of the observer's visual field and is
covered with a random pattern of colored dots, containing no clues about gravitational orientation . The centre of the dome is fixed to
the shaft of a DC torque motor; a circular target of 14-deg visual angle with a straight line through the centre and mounted on a coaxial shaft connected to the DC servomotor is 30 cm in front of the observer. The patient has to adjust the central test edge to the vertical, using a potentiometer. The output of the potentiometer is automatically recorded in degrees on a Pe. SVV is determined by
calculating the means of 10 adjustments of the target disk from a random offset position to the subjective vertical. Under these conditions, the normal range (x 2 sD) of the SVV is 2.5 degs. As apart of the ocular tilt reaction, acute lower medullary lesions (e.g. in
Wallenberg's syndrome) and unilateral peripheral vestibular lesions cause an ipsiversive displacement, whereas mesencephalic lesions
cause a contraversive displacement of the SW.

46

Vertigo

( day 3 )

( day 30 )

Fig.2.22. Measurement of the oeu/ar torsion (OT, cyclorotation) bya scanning laser ophthalmoscope (SLO); top. Photographs can be
made with the SLO without requiring administration of a mydriatic drug. OT is measured with the patient sitting and the head upright.
Photographs are taken for both eyes separately during fixation of a central target; an example is shown (bottom): fundus of the left eye
of a patient with a left-sided vestibular neuritis on days 3 and 30 after symptom onset. The position of the eye in the roll plane is
measured as the angle formed bya straight line through the papilla and fovea (papilla-fovea meridian) and a horizontalline.
According to this method, both eyes of healthy controls had a slightly excyclotropic position in the roll plane (i.e. counterclockwise
rotation of the right eye, clockwise rotation of the the left eye, from the viewpoint of the examiner). The normal range of OT (mean
2 SD) is -1 to 11.5 degs. As the displacement of the subjective visual vertical (see Fig. 2.21). OT is also a hallmark of the ocular tilt
reaction (see Fig. 2.5); a binocu/ar ipsilateral cyclorotation is observed in acute lower medullary lesions (e.g. in Wallenberg's syndrome)
and unilateral peripheral vestibular lesions, whereas a contraversive cyclorotation is seen in mesencephalic lesions.

Approaching the patient

References
Andermann F, Cosgrove JBR, Lloyd-Smith D, Walters AM (1959)
Paroxysmal dysarthria and ataxia in multiple sclerosis.
Neurology (Minneap) 9:211-215
Aschan G, Bergstedt M, Goldberg L, Laurell L (1956) Positional
nystagmus in man during and after alcohol intoxication. J Stud
AlcohoI17:381-405
Auramo Y, Juhola M, Pyykk I (1993) An expert system for the
computer-aided diagnosis of dizziness and vertigo. Med Inform
18:293-305
Baloh RW (1996) History. I. Patient with dizziness. In: Baloh RW,
Halmagyi GM (eds) Disorders of the vestibular system. Oxford
University Press, NewYork, pp 157-170
Baloh RW, Jacobsen K, Honrubia V (1993) Horizontal semieircular
canal variant of benign positional vertigo. Neurology 43:
2542-2549
Barany R (1911) Experimentelle Alkoholintoxikation. Monatschr
Ohrenheilk 45:959-962
Bhmer A, Straumann D, Fetter M (1997) Three-dimensional
analysis of spontaneous nystagmus in peripheral vestibular
lesions. Ann Otol Rhinol Laryngol106:61-68
Brandt Th (1997) Cortical matching ofvisual and vestibular 3D
co ordinate maps. Ann NeuroI42:983-984
Brandt T, Daroff RB (1980) The multisensory physiological and
pathological vertigo syndromes. Ann NeuroI7:195-203
Brandt Th, Steddin S (1993) Current view of the mechanism of
benign paroxysmal positioning vertigo: cupololithiasis or
canalolithiasis? J Vestib Res 3:373-382
Brandt Th, Dieterich M (1994a) Vestibular paroxysmia. Neuroophthalmology 14:359-369
Brandt Th, Dieterich M (1994b) Vestibular paroxysmia: vascular
compression of the eighth nerve? Lancet i:798-799
Brandt T, Strupp M (1997) Episodic ataxia type 1 and 2 (Familial
periodic ataxia/vertigo). Audiol NeurootoI2:373-383
Brevern M von, Faldon ME, Brookes GB, Gresty MA (1997)
Evaluating 3D semieircular canal function by perception of
rotation. Am J OtoI18:484-493
Brunt ER, van Weerden TW (1990) Familial paroxysmal kinesigenic ataxia and continuous myokymia. Brain 113:1361-1382
Bucher SF, Dieterich M, Wiesmann, M, Weiss A, Zink R, Yousry TA,
Brandt T (1998) Cerebral functional MRI of vestibular, auditory
and nociceptive areas during galvanic stimulation. Ann Neurol
44:120-125
Bttner U, Straube A, Brandt Th (1987) Paroxysmal spontaneous
nystagmus and vertigo evoked by lateral eye position.
Neurology 37:1553-1555
Cox TA, Corbett J], Thompson HS, Lennarson L (1981) Upbeat
nystagmus changing to downbeat nystagmus with convergence.
Neurology 31:801
Cremer PD, Halmagyi GM, Aw ST, Curthoys IS, McGarvie LA,
Todd MJ, Black RA, Hannigau IP (1998) Semicircular canal
plane head impulses detect absent function of individual semieircular canals. Brain 121: 699-716
Deecke L, Mergner T, Plester D (1981) Tullio phenomenon with
torsion of the eye and subjective tilt of the visual surround. Ann
NY Acad Sei 374:650-655
Dieterich M, Brandt T, Fries W (1989) Otolith function in man:
results from a case of otolith Tullio phenomenon. Brain
112:1377-1392
Espir MLE, Millac P (1970) Treatment of paroxysmal disorders in
multiple sclerosis with carbamazepine (Tegretol). J Neurol
Neurosurg Psychiatry 33:528-531
Fetter M, Haslwanter T, Misslich H, Tweed D (1997) Threedimensional kinematics of eye, head and limb movements.
Harwood Academic Publishers, Amsterdam
Foerster 0 (1936) Sensible corticale Felder. In: Bumke 0, Foerster

47

o (eds) Handbuch der Neurologie, Vol VI, Springer, Berlin,


pp 358-448
Fries W, Dieterich M, Brandt T (1993) Otolith contributions
to postural control in man: short latency motor responses
following sound stimulation in a case of otolith Tullio phenomenon. Gait&Posture 1:145-153
Gresty MA, Bronstein AM, Brandt T, Dieterich M (1992)
Neurology of otolith function. Peripheral and central dis orders.
Brain 115:647-673
Griggs RC, Nutt JG (1995) Episodic ataxias as channelopathies.
Ann NeuroI37:285-287
Halmagyi GM, Curthoys IS (1988) A clinical sign of canal paresis.
Arch NeuroI45:737-739
Haynes BF, Kaiser-Kupfer MI, Mason P, Franci AS (1980) Cogan
syndrome: studies in 13 patients, long-term follow-up, and a
review of the literature. Medieine 59:426-441
Hedges TR, Hoyt WF (1982) Ocular tilt reaction due to an upper
brainstem lesion. Paroxysmal skew deviation, torsion,
and oscillation of the eyes with head tilt. Ann Neurol
11:537-540
Honrubia V (1994) Quantative vestibular function tests and the
clinical examination. In: Herdman SJ (ed) Vestibular rehabilitation. Davis, Philadelphia, pp 113-164
Jannetta PJ, Moller MB, Moller AR (1984) Disabling positional
vertigo. N Engl J Med 310: 1700-1705
Jongkees LB, Maas JP, Philipszoon AJ (1962) Clinical nystagmography: a detailed study of electro-nystagmography in 341
patients with vertigo. Pract Otorhinolaryngol 24:65-93
Laicher S, Linzer M (1996) Syncope. In: Brandt Th, Caplan LR,
Dichgans J, Diener H-C (eds) Neurological dis orders. Course
and treatment. Academic Press, San Diego, pp 1063-1067
Leigh RJ, Zee DS (1991) The neurology of eye movements. Davis,
Philadelphia
Lueck CI, Hamlyn P, Crawford TI, Levy IS, Brindley ES, Watkins
ES, Kennard C (1991) A case of ocular tilt reaction and torsion al
nystagmus due to direct stimulation of the midbrain in man.
Brain 114:2069-2079
Mira E, Buizza A, Magenes G, Manfrin M, Schmid R (1990) Expert
systems as a diagnostic aid in otoneurology. ORL 52:96-103
Moller MB, Moller AR, Jannetta PJ, Sekhar L (1986) Diagnosis and
surgical treatment of disabling positional vertigo. J Neurosurg
64:21-28
Money KE, Myles WS, Hoffert BM (1974) The mechanism of positional alcohol nystagmus. Can J OtolaryngoI3:302-313
O'Connor KP, Hallam RS, Hinchcliffe R (1989) Evaluation of a
computer interview system for use with neuro-otology patients.
Clin OtolaryngoI14:3-9
Olivia A, Rosenberg ML (1990) Convergence-evoked nystagmus.
Neurology 40:161-162
Osterman PO, Westerberg CE (1975) Paroxysmal attacks in multiple
sclerosis. Brain 98:189-202
Penfield W, Jasper H (1954) Epilepsy and the functional anatomy
of the human brain. Little Brown, Boston
Rabinovitch HE, Sharpe JA, Sylvester TO (1977) The ocular tilt
reaction. A paroxysmal dyskinesia associated with elliptical
nystagmus. Arch Opthalmol 95: 1395-1398
Ropper HA (1983) Illusions of tilting of the vestibular environment: report of 5 cases. J Clin NeuroophthalmoI3:147-151
Schneider RC, Calhoun HD, Crosby EC (1968) Vertigo and rotational movement in cortical and subcorticallesions. J Neurol
Sei 6:493-516
Sharpe JA, Hoyt WF, Rosenberg MA (1975) Convergence-evoked
nystagmus. Congenital and acquired forms. Arch Neurol 32: 191
Stodieck SRG, Brandt Th, Bttner U (1990) Visual and vestibular
epileptic seizures. Electroenceph Clin Neurophysiol 75:65-66P
Stolz SE, Chartrian G-E, Spence AM (1991) Epileptic nystagmus.
Epilepsia 32:910-918
Straumann D, Zee DS (1995) Three-dimensional aspects of eye
movements. Curr Opin NeuroI8:69-71

48
Tiliket L, Ventre-Dominey 1, Vighetto A, Grochowicki M (1996)
Room tilt illusion: a central otolith dysfunction. Arch Neurol
53:1259-1264
Tullio P (1929) Das Ohr und die Entstehung der Sprache und
Schrift. Urban and Schwarzenberg, Munich.
Tusa RJ, Kaplan PW, Hain TC, Naidu S (1990) Ipsiversive eye deviation and epileptic nystagmus. Neurology 40:662~665
Zee DS (1978) Ophthalmoscopy in examination of patients with
vestibular dis orders. Ann NeuroI3:373-374

Vertigo
Zink R, Steddin S, Weiss A, Brandt Th, Dieterich M (1997)
Galvanic stimulation in humans: effects on otolith function in
roll. Neurosci Lett 232:171-174
Zink R, Bucher SF, Weiss A, Brandt Th, Dieterich M (1998) Effects
of galvanic vestibular stimulation on otolithic and semicircular
canal eye movements and perceived vertical. Electroenceph
Clin Neurophysiol107:200-205

Management of the dizzy patient

The prevailing good prognosis of vertigo should be


emphasised, because

many forms of vertigo have a benign cause and


are characterised by spontaneous recovery of
vestibular function or central compensation of a
peripheral vestibular tone imbalancej
most forms of vertigo can be effectively relieved
by pharmacological treatment (Table 3.1), physical therapy (Table 3.2), surgery (Iable 3.3), or
psychotherapy (Chaps. 31 and 32).

There is, however, no common treatment, and


vestibular suppressants (Iable 3.4) provide only
symptomatic relief of vertigo and nausea. A specific
therapeutic approach thus requires recognition of
the numerous particular pathomechanisms involved.
Such therapy can include causative, symptomatic, or
preventive approaches; all are discussed in detail in
the chapters describing the particular disorder.
lable 3.1. Pharmacological therapies for vertigo
Therapy

Vertigo

Vestibular suppressants

Symptomatic relief of nausea (in acute


peripheral and vestibular nudei lesionsJ,
prevention of motion sickness

Antiepileptic drugs

Vestibular epilepsy, vestibular paroxysmia


(disabling positional vertigoJ, paroxysmal
dysarthria and ataxia in MS, other central
vestibular paroxysms, superior oblique
myokymia

Beta-receptor blockers

Basilar migraine (benign recurrent vertigo)

Betahistine

Meniere's disease

Antibiotics

Infections of the ear and temporal bone

Ototoxie antibiotics

Meniere's disease (Meniere's drop attacks)

Corticosteroids

Vestibular neuritis, autoimmune inner ear


disease

Badofen

Downbeat or upbeat nystagmus or vertigo

Acetazolamide

Familial periodic ataxia or vertigo

Table 3.2. Physical therapies for vertigo


Therapy

Vertigo

Benign paroxysmal positioning vertigo


Vestibular rehabilitation, central
compensation of acute vestibular loss,
habituation for prevention of motion
sickness, improvement of balance skills
(e.g. in the elderly)
Physical therapy (neck collar) Cervicill vertigo (?)

Deliberate manoeuvres
Vestibular exercises

From Brandt (1993).

Table 3.3. Surgical interventions for vertigo


Surgery

Vertigo

Surgical decompression of
eighth nerve
Surgieal decompression of
vertebral artery
Ampullary nerve seetion or
canal plugging
Endolymphiltic shunt
Vestibular nerve section or
labyrinthectomy
Neurovascular decompression?

Tumour (acoustic neurinoma) or eyst

Surgical patching

Rotational vertebral artery occlusion


Benign paroxysmal positioning vertigo
Meniere's disease
Intractable Meniere's disease
Vestibular paroxysmia (disabling
position al vertigo)
Perilymph fistula

From Brand! (1993).

Antivertiginous and antiemetic drugs


A variety of drugs used for symptomatic relief of
vertigo and nausea have the major side effect of general sedation (Brandt et al. 1974j Foster and Baloh
1996). Vestibular suppressants, including anticholinergics, antihistamines, and benzodiazepines, provide
symptomatic relief of distressing symptoms by
downregulating vestibular excitability. Antiemetics
preferably control nausea and vomiting by acting on
the medullary vomiting centre, the chemoreceptor

From Brandt (1993).

49

Vertigo

50

trigger zone, or the gastrointestinal tract itself.


Vestibular suppressants are often acetylcholine and
histamine antagonists, which act as acetylcholine
antagonists by competitive inhibition at muscarinic
receptors in the vestibular nuclei, their most
likely site of action. Vestibular suppression by
benzodiazepines is best explained by their GABA A
agonistic effect, because GABA is the major neuroinhibitory transmitter for vestibular neurons.
Antiemetics are effective mainly due to their
dopamine (D z) antagonist properties, but some
antiemetics also have muscarinergic or antihistaminic (H j ) properties that may assist in vestibular
suppression as weIl. Primary vestibular suppressants
such as scopolamine also effectively suppress vomiting by virtue of their muscarinergic action.
Antiemetics are more selective in action. They are
primarily used to control nausea and vomiting; for
treatment of severe vertigo with nausea, they are
often combined with antivertiginous drugs (Foster
and Baloh 1996).
There are only four clear indications for the use of
antivertiginous (vestibular suppressants) and
antiemetic drugs to control vertigo, nausea, and
vomiting (Brandt 1993):
to prevent nausea due to acute peripheral
vestibulopathy (for the first 1-3 days or as long
as nausea lasts),
2. to prevent severe vertigo and nausea due to acute
brainstem or archicerebellar lesions near the
vestibular nuclei,
3. to prevent severe vertigo attacks recurring on a
frequent basis, and
4. to prevent motion sickness.
1.

For conditions 1 and 2, fast-acting compounds with


vestibular and general sedation should be preferably
administered, e.g. diazepam or promethazine combined with dimenhydrinate if nausea and vomiting
are exceptionally severe. These drugs should not be
given after nausea has disappeared, because they
prolong the time course of central compensation of
an acute vestibular tone imbalance.
Mobility and vestibular excitability are major
requirements for recovery and vestibular rehabilitation. Readjustment of the vestibular reflexes, which
act on eye and body muscles, requires sensory feedback from the sensory mismatch elicited by voluntary movements. Therefore, on the basis of our
current knowledge of vestibular physiology, continued management should consist of vestibular exereises that promote central compensation (Table 4.1).
Antivertiginous and antiemetic drugs are not indicated for patients suffering from chronic dizziness. A
prophylactic treatment with vestibular suppressants,

Table 3.4.

Commonly used antivertiginous and antiemetic drugs

Drug

Dosage

Anticholinergics
Scopolamine
0.6 mg po q 4-6 h or
(Transderm Scop) Transdermal patch:
1 q 3 days
Antihistamines
Dimenhydrinate
(Drama mine)

Meclizine

50 mg po q 4-6 h or
im q 4-6 h or
100 mg suppository
q 8-10 h
25 mg po q 4-6 h

(Antivert, Bonine)
Promethazine
15 or 50 mg po q 4-6h or
(Phenergan)

im q 4-6 h or
suppository q 4-6 h

Phenothiazine
Prochlorperazine 5 or 10 mg po q 4-6 h or
(Compazine)
im q 6 h or
25 mg suppository q 12 h

Action
Muscarine antagonist

Histamine (H,)
antagonist
Muscarine antagonist

Histamine (H,)
antagonist
Muscarine antagonist
Histamine (H,)
antagonist
Muscarine antagonist
Dopamine (D 2)
antagonist
Muscarine antagonist
Dopamine (D 2)
antagonist

Butyrophenone
Droperidol
(Inapsine)

2.5 or 5 mg im q 12 h

Muscarine antagonist
Dopamine (D 2)
antagonist

Benzodiazepines
Diazepam
(Valium)
Clonazepam
(Klonopin)

5 or 10 mg po bid-qid
im q 4-6 h or iv q 4-6h
0.5 mg po tid

GABA A agonist
GABA A agonist

e.g. scopolamine or dimenhydrinate, is justified only


in exceptional situations of rare patients who have
frequent and severe vertigo attacks. In severe cases
ofbenign paroxysmal positioning vertigo (p. 265) it
may become necessary to control nausea and
vomiting when performing physical liberatory
manoeuvres. It is our own experience that severe
central positioning vomiting (p. 293) is best controlled by benzodiazepines rather than antiemetics
or typical vestibular suppressants. Scopolamine
administered transdermally as Transderm Scop provides a continuous blood level over a 3-day period
and effectively prevents motion sickness. The selection of vestibular suppressants and antiemetic drugs
should take into account that those that reach a peak
effect 7-9 hours after ingestion (Manning et al. 1992)
are ineffective for treating short vertigo attacks.
Other effective drugs can be expected to be developed from compounds that interfere with the presynaptic histamine receptor H 3 (Timmerman 1994;
Kingma et al. 1997) or GABAA receptors (Ehrenberger
and Felix 1996). Some reliable pharmacological therapies are available for abnormal vestibular and non-

Management ofthe dizzy patient


vestibular eye movements, which override fixation
and thus cause oscillopsia and impair vision (Leigh
et al. 1994). The GABA B agonist baclofen suppresses
periodic alternating nystagmus in patients (Halmagyi
et al. 1980) and animals with experimentallesions of
the nodulus and uvula (Waespe et al. 1985). The
GABAergic anticonvulsant gabapentine (AverbuchHeller et al. 1997) and the glutamate antagonist
memantine (Starck et al. 1997) also effectively suppress acquired pendular nystagmus. Baclofen provides an effective treatment of some patients with
downbeat or upbeat nystagmus (Dieterich et al.
1991); occasional patients will also respond to
gabapentine (Averbuch-Heller et al. 1997).

Surgical treatment
Surgical procedures for the treatment of dizzy
patients are primarily used by otolaryngologists but
also to a minor extent by neurosurgeons. There is no
doubt that surgery is the therapy of first choice, e.g.
for acoustic neurinomas or an infratentorial cavernoma. The same holds for the rotational vertebral
artery syndrome (p. 296), because of the danger of
vertebral artery occlusion or embolism. In these
cases, vertigo may be part of the clinical syndrome,
but the indication for surgery is based mainly on the
impending risk of brain and cranial nerve damage.
Indications for surgical interventions based only on
the goal to control recurrent or chronic vertigo are
rare and should always be considered second choice
after conservative management has failed. The multiple procedures can be classified as

non-destructive
decompression of the eighth nerve (acoustic
neurinoma, cerebellopontine angle cyst)
neurovascular decompression of the eighth
netve (vestibular paroxysmia)
endolymphatic shunt in Meniere's disease
surgical patching of perilymph fistulas
selectively destructive
retrolabyrinthine or middle fossa vestibular
nerve section in intractable Meniere's disease
semicircular canal plugging or ampullary nerve
section in intractable benign paroxysmal positioning vertigo
destructive
oval window or trans mastoid labyrinthectomy
translabyrinthine vestibular nerve section
laser labyrinthectomy

Surgery is still most often considered for treatment of Meniere's disease (p. 92). Thomsen et al.

51

(1981) have criticised endolymphatic sac surgery,


since they found no statistical difference between
endolymphatic sac. surgery and placebo surgery.
However, their study has also been criticised in turn
for its small sampie size and problems with design
and statistical methods (Vaisrub 1981). Many otolaryngologists continue to believe in the beneficial
effect of this relatively safe, non-destructive procedure (Brackmann 1996). Vestibular nerve section is
still performed in rare patients with disabling vertigo and preserved hearing in whom medical treatment and endolymphatic sac shunts have failed.
Vestibular neurectomy poses risks of facial paresis,
meningitis, cerebral fluid leakage, or epidural
haematoma (Gacek and Gacek 1996). Labyrinthectomy of the affected ear can be performed only in
cases with associated intractable hearing loss. Both
procedures, vestibular neurectomy and labyrinthectomy, were found to be equally effective in relieving
vertigo (Gacek and Gacek 1996), although deafferentation of the vestibular end-organ is sometimes
incomplete, as shown for bilateral vestibular neurectomy (Bhmer and Fisch 1993). Laser labyrinthectomy
has also been performed in animals and patients
(Nomura et al. 1993), with the aim of selectively
destroying individual otolithic or semicircular canal
structures. Despite vestibular compensation, vestibular neurectomy results in permanent dynamic
vestibulo-ocular reflex deficits (Halmagyi et al. 1991;
Kanayama et al. 1995). A considerable percentage of
patients will never sufficiently compensate and will
continue to suffer from chronic dizziness and disequilibrium (Halmagyi 1994).
According to most available reports, the less effective results of destructive surgery for non-Menihe's
vertigo (Benecke 1994) are difficult to evaluate, since
convincing analysis of the specific diagnoses is lacking (Kemink et al. 1991). These operations were
based on diagnoses such as "chronic vestibular
neuronitis" (Benecke 1994) or "uncompensated
vestibular neuritis" (Kemink et al. 1991). Clinical
experience has taught us that many of these socalled chronic conditions actually represent the
transition of a peripheral vestibular dysfunction to a
psychosomatic disease, e.g. phobic postural vertigo
(p.469).
Transtympanic aminoglycoside treatment of
Meniere's disease (Bergenius and dkvist 1996) also
offers control of vertigo at the risk of profound hearing 10ss.The aminoglycoside treatment is increasingly
being preferred to surgery (p. 91). Surgical procedures for benign paroxysmal positioning vertigo
(Pohl 1996) include either transmeatal singular
neurectomy or semicircular canal plugging (p. 269).
If physical liberatory manoeuvres are performed
correctly and continuously over a sufficiently long

52

time, it is our experience that these procedures are


required for only exceptional patients.

Vestibular exercises and physical


therapy tor vestibular rehabilitation
Vestibular exercises are performed either to promote
central habituation so as to prevent motion sickness
(see motion sickness; p. 491) or to readjust vestibuloocular and vestibulospinal reflexes as a form of
retraining for exceptional populations (see vestibular compensation; p. 76).
Animal experiments have shown that exercise
may facilitate vestibular compensation (Lacour et al.
1976; Courjon et al. 1977; Igarashi 1986; Fetter and
Zee 1988). The special role of visual input has been
convincingly demonstrated by Courjon and
Jeannerod (1979) and Lacour and Xerri (1981).
Furthermore, animal experiments suggest that there
is a critical period for functional recovery which is
crucial for achieving either optimal or minimal
repair (Xerri and Lacour 1980; Lacour 1984). The few
available clinical control studies provide evidence
that physical therapy is superior to general conditioning exercises. Such physical therapy helps
patients with chronic dizziness (Horak et al. 1992)
and those after resection of an acoustic neuroma
recover balance earlier than if not treated (Herdman
et al. 1995); similarly patients after acute unilateral
vestibular neuritis exhibit normalisation of postural
sway within a significantly shorter time course than
the control group (Strupp et al. 1998b).
Vestibular compensation is no "simple or single"
process. It consists of multiple processes for perceptual, vestibulo-ocular, and vestibulospinal readjustment, which have different time courses at different
sites in the brain and the spinal cord (p. 56).
Therefore, vestibular rehabilitation should incorporate different exercises that involve eye, head, and
body movements and the monitoring of patients'
progress separately for the different perceptual,
ocular motor, and postural vestibular functions. A
study on the efficacy of vestibular exercises for compensation and substitution following an acute unilateral partial vestibular loss (cases of vestibular
neuritis without recovery of peripheral function
during the training phase) found that only postural
balance was significandy facilitated (Fig. 4.6), not the
time course of recovery of ocular torsion or tilts of
perceived vertical as measured in degrees (Strupp et
al. 1998b). Recovery from bilaterallabyrinthine loss
was also demonstrated in animal experiments

Vertigo
(Igarashi et al. 1988) and in patients with chronic
bilateral vestibular deficits (Krebs et al. 1993; Szturm
et al. 1994). In such cases recovery takes place more
slowly and incompletely, leaving permanent instability during intensified balance tasks and in darkness
as weIl as rapid head movements while walking. Up
to now no controlled studies have focused on the
effects of physical exercise on the rehabilitation of
patients with central vestibular dis orders. However,
the patients' rapid recovery, e.g. from lateropulsion
in Wallenberg's syndrome, when they become
mobilised and are able to perform intensive physical
therapy seems to support the efficacy of exercises,
but provides no proof.

Quantitative effects of balance training on


postural sway in normal subjects
The remarkable balancing skills of steeplejacks,
tight-rope artists, and gymnasts indicate that postural control is not optimised under daily conditions, but can be greatly improved by training. It has
been shown that central nervous system plasticity
can compensate for peripheral or central neurological deficits by sensorimotor rearrangement, and this
is the rationale of some physical therapy approaches
to rehabilitating patients with acquired vestibular
ataxias (Brandt et al. 1986; Brandt and Paulus 1989).
There are a number of important issues of concern
to persons involved in retraining balance in patients
with postural instability. Questions to be answered
include the following: what are the short- and longterm effects of training on postural instability? What
are the hypothesised neural mechanisms underlying
training effects? Wh at experimental evidence is
there to support the existence of an ability to train
balance by stimulation of different sensory systems?
Finally, is there evidence to suggest that different
training methods may not all be equally effective?
The effects of postural training on performance
in a balancing task were investigated in healthy control subjects to assess the likely utility of training for
ataxic patients. Subjects were artificially destabilised,
either by standing with the head maximally extended
or by standing on one foot, and the time course of
any improvement in performance was monitored
(Brandt et al. 1981; Bchele et al. 1984). It can be
demonstrated that learning by training clearly correlates with the degree of postural instability: the
greater the initial risk of falling, the greater the percent reduction in sway amplitudes produced by
training. Much insight can be gained by examining
effects of training on experimentally induced postural instability in healthy subjects. Results from
these experiments can be used to design effective

Management of the dizzy patient

53

therapies for re training equilibrium in patients with


peripheral or central nervous system pathologies.
Symptoms of to-and-fro vertigo and postural
imbalance occur frequently in healthy people during
head extension (see head extension and vertigo;
p. 297). Posturographic measurements reveal a significant increase in amplitudes of body sway
induced by head extension alone, particularly when
the "stabilising" input of the non-vestibular sensory
systems is reduced or eliminated, for example when
the eyes are closed while the subject is standing on
foam rubber (Fig. 18.4). Alterations in the reafferent
vestibular consequences of self-generated body sway
may be responsible for the impairment of postural
control in the offending head position (p. 297;
Brandt et al. 1981). With the eyes open, there is no
significant increase in sway amplitudes. This is
indicative of the redundant, multiloop control of
posture with overlapping functional ranges of the
stabilising systems, enabling them to compensate
partially for each others' deficiencies. One hour of
intermi ttent practice (total time of head extension,
15 min) produced a remarkable improvement ofbalance. The mean reduction in sway amplitudes for all
subjects was 20-30% RMS (root-mean square) value
during the first trial (Figs 3.1 and 3.2). This represents an exponential short-term training effect.

Figures 3.1 and 3.2 show that a "short-term daily


training effect" and a "long-term training effect"
together form a characteristic saw-tooth-like curve
of sway activity, which within 5 days reaches an
asymptote at 40-50% of the initial sway activity
(Brandt et al. 1981). After training ceases, control
measurements reveal that learned balance skill without continued practice gradually disappears, even if
it does not return to the initial pretraining values for
lateral body sway within 40 days (Fig. 3.3). Other
experience teaches us that the lack of sensorimotor
exercise causes a decrement in balance performance,
which is best known as transient ataxia after bedrest
(Taylor et al. 1949; Haines 1974). Balance impairment after prolonged bedrest, which improves within 1 to 3 recovery days, must be independent of
muscle strength, because it cannot be avoided by
performing isotonic or isometrie exercises in bed.

Balance training in vestibular disorders


What do the effects of balance training on postural
sway in normal subjects imply for clinical application? The differential effects of balance training on
postural sway with the eyes open and the eyes closed
clearly show that the process of sensorimotor

5th day

1st day
post.... al sway

head
extension

head
extension

l
...J..Q!...,

J~
a
r

J~
a

]~

after 1h

after 1h

f-----i

1-----1

Fig.3.1. Effects of balance training on body sway. Fore-aft and lateral body sway of anormal subject (original recordings) with the
eyes closed and head extension at the beginning a, after I hof intermittent practice b, at the beginning of the fifth c and at the end of
the fifth training day d. A short-term training effect is evident in the comparison of a and b, and c and d. A long-term training effect is
evident in the comparison of a and c and band d. (From Brandt et al. 1981.)

54

Vertigo
fore - aft body sway

lateral body sway

RMS

100%

100%

.,
:.

.,5
c

e'"
GI
0

Ci

50

:~'"

eyes c losed

'~~

~~v

''""
'.,"
~

eyes closed

~~~

~~"-<>-o~~

eyes open

eyes open

0-<>-...0

n 20
0

5 day

50

Fig.3.2. Postural balance with head extension. Effects of training over 5 days. Percentage of reduction in mean root-mean square
values of fore-aft and lateral body sway for 20 subjects during the course of a training period of 5 days (1 h per day: total time of head
elevation with the eyes open, 8 min; eyes closed, 8 min). A "short-term daily training effect" (represented by the three measurements
made during 1 h) and a "Iong-term training effect" form a characteristic saw-tooth-like curve of sway activity, which reaches an asymptote at 40-50% of the initial sway activity. No significant training effect can be recognised for the condition with eyes open in which
the initial instability was less than with eyes closed. (From Brandt et al. 1981.)

..

fOl' e-lIIIh body SwA)'

000

II !
L

90

.,

80

Ir

10
60
50

end 01 traInIno

40 ""YI

90

II !
~

80

Ir

30

Lateral body sw8y

'00

.,

20

'0

10
80

50

end 01 tr.'ning

'0

20

30

Fig.3.3. Postural balance with head extension; persisting training effects. The duration of the effect of training on postural balance with head extension and eyes closed for 10 subjects.
Control measurements of body sway after a training period of 5
days reveal that without ongoing practice, postural imbalance
increases but only for fore-aft activity, and tends to reach the pretraining values within 40 days. (From Brandt et al. 1981 .)

rearrangement with subsequent postural stability is


related to the degree of the initial instability.
Consequently, the clinician should devise a strategy
of ataxia therapy which exposes the patient increasingly often to unstable body positions in order to
facilitate rearrangement and recruitment of control
capacities. Not only is the rapid improvement striking, in spite of the relatively short training phases,
but also the long duration of the newly acquired
recalibration of sensorimotor controlloops, which
amounts to many weeks. Stance and gait aids seem
to improve stability, but they only transiently alleviate patients' balance problems. When used continuously, they worsen the symptoms, because the
multiloop control rapidly adapts to the additional
feedback and support that the aids provide. Patients
with acquired downbeat nystagmus (p. 201), for
example, who suffer from a tendency to fall backward, improve their postural stability during balance
training despite the persistence of the nystagmus.
Recovery may require only a few days or as long as
several months. In the particular clinical case, however, it is not clear whether this benefit from retraining arises from functional recovery of lesioned
brainstem structures or central compensation by
sensorimotor re arrangement.
The goals of physical therapy and intervention are
to improve the patient's functional balance and
mobility, general physical condition and activity
level, and safety during ambulation and other gaitrelated activities, and to decrease the magnitude of

Management of the dizzy patient

distressing disorientation (Herdman 1996). Physical


therapy evaluation should
1. identify the particular problem of each patient,
2. establish the goals of treatment, and
3. monitor the patient's progress during treatment
(Herdman 1996).

Exercises for patients with acute vestibular dysfunction consist mainly of eye, head, and body
movements designed to provoke a sensory mismatch; they enhance compensation by facilitating
central recalibration, although the symptoms are initially uncomfortable.
Various types of artificial feedback have been
~se~ for retraining balance in patients with disequihbnum. Postural strategies are often characterised
~n terms of feedback interaction between sensory
mput and motor output. In fact, reduction of body
sway may be seen as an example of feedforward as
weIl as feedback strategies: the body has to detect
and to reduce its own self-generated sway.
Feedforward strategies in posture and locomotion
are used to adjust the body to expected positions in
space; feedback strategies react to sensed deviations
from the ideal antigravity position (Sollwert). This
"natural" feedback is usually achieved unconsciously,
so that posture can be stabilised despite full concentration on separate voluntary goal-directed
limb movements. Although several authors have
~escribed the possibility of decreasing body oscillahon by using artificial visual, acoustic, or
somatosensory feedback, the measurable effects are
insignificant. Thus, their use for physical therapy of
ataxias appears limited (Brandt and Paulus 1989).

55

recovery with the disappearance of nystagmus and


ataxia following unilateral labyrinthine loss.
Vestibular compensation, the central nervous system's capacity for plastic adaptive change, has
become a basic paradigm in the study of lesioninduced neuronal plasticity (Flohr et al. 1981).

Terms and definitions of plasticity and central


compensation
There is considerable terminological confusion in
the description of phenomena of plasticity at biochemical, electrophysiological, structural, functional, or behavioural levels (Table 3.5). Not only are
many of the terms imprecisely defined, but their
definitions have changed over the last decades.
P.lasticity, for example, is used to refer to changes
eher at the cellular/network level (neural plasticity)
or at the behaviourallevel (behavioural plasticity).
The ~nderlying mechanisms of neural plasticity are
multIple; they include biochemical, structural, and
functional changes. The consequences of these
changes, which express themselves in behavioural
plasticity, are likewise multiple. A short list of commonly used key words for both neural and behavioural aspects of plasticity is presented in Table 3.5.
In most cases the causal correlation between neural
and behavioural plasticity is not known. Some terms
are used as synonyms, such as restitution and
restoration. Recovery is a largely descriptive term,
which is related to function or behaviour, independently of the underlying mechanism or structural
changes. The list of terms in tabular form reveals the
Table 3.5.

Plasticity of the vestibular system:


central compensation and sensory
substitution for vestibular deficits
Acute unilateral vestibular lesions produce a tone
imbalance characterised by direction-specific spontaneous nystagmus and unstable posture. Initially,
the vestibulocerebellum seems to reduce somewhat
the distressing consequences of this tone imbalance
by decreasing the sensitivity of vestibular nuclei
neurons. However, these symptoms only abate when
the tone is re-equalised, thereby restoring peripheral
functions by central compensation and multisensory
substitution. "Active repair" is provided by slow central vestibular compensation, wh ich normally takes
days to some weeks. Early reports of Bechterew
(1883) and Ewald (1892) described a gradual

Words and terms describing plasticity at neural and


behaviourallevels'
Neural plasticity

Behavioural plasticity

Facilitation
Synaptic depression
Synaptic potentiation
Receptor upregulation
Receptor downregulation
Sprouting
Supersensitivity

Adjustment
Compensation
Habituation
Learning
Preprogramming
Readjustment
Recalibration
Recovery
Rehabilitation
Restitution
Restoration
Selection
Sensitisation
Substitution
Strategy

'These terms are often poorly defined, and some are used to describe
neural as weil as behavioural plasticity.
From Brandt et al. (1997).

Vertigo

56

basic dichotomy of plastic phenomena at neural and


behaviourallevels.
This discussion of plasticity of the vestibular
system focuses on four of the most relevant terms:
(1) adaptation, (2) habituation, (3) sensitisation, and
(4) compensation (Table 3.6). Adaptation means the
adjustment of a sensory system to its environment
or the process by which this ability is achieved.
Sensory receptor adaptation is thought to be an
important component of perceptual adaptation. For
instance, rapidly adapting mechanoreceptors, such
as the pacinian corpuscle, respond transiently only at
the onset and end of a change in stimulus position.
With respect to the vestibulo-ocular reflex (VOR), a
reflex controlled in a "feedforward" or "open loop"
manner, short- and long-term adaptations are bidirectional. The VOR can adapt from fixating distant
objects to fixating near objects. For the image of a
distant object to remain on the fovea of the retina
during a head rotation, eye rotation of equal velo city
but opposite direction must be generated. When this
perfect compensation is achieved, the gain of the
response (eye movement/head movement) is 1.0.
This gain must be larger than 1.0 when we fixate
near objects binocularly because of the different
axes of rotation for eyes and head. Another example:
the retinal image in a person who wears glasses for
myopia is smaller than that ordinarily projected by
the lens in his eye. When he moves his head, the
image movement is less; to stabilise the image the
eyes must also move less. In such a case the gain is
less than 1.0. The VOR even adapts to inverted
prisms (Gonshor and Melvill Jones 1976), i.e. subjects who have worn reversing prisms for several
days move their eyes in the same direction as the
head, even in darkness.
Habituation, the simplest form of learning, is
defined in Dorland's Illustrated Medical Dictionary
Table 3.6. Direction of changes during adaptation, habituation, compensation, substitution and recovery
Direction of change

ADAPTATION

tt

HABITUATION

tt

COMPENSATION

SUBSTITUTION

f- ---7

RECOVERY

f- ---7

From Brandt et al. (1997).

Example
Changes in ga in of vestibulo-ocular
reflex (VOR) (induced by convergence
or inverted prisms)
Motion sickness, motion perception
(decrement in perceived velocity
during prolonged stimulation)
Complex recovery after peripheral
unilateral vestibular 1055
Vestibular by visual or somatosensory
input, ocular slow phases by saccades
(defective VOR)
Complex functional repair after alesion

(1994) as "the gradual adaptation to a stimulus or


the environment" (note that frequently one term is
explained by another term). A more precise definition is given by Thompson and Spencer (1966):
habituation is a central process that is independent
of sensory adaptation and motor fatigue. This is best
reflected by the apparent decrement in perceived
velo city during a prolonged car ride or the habituation to motion sickness on a ship in a rough sea
within 3 days (p. 487). The diagram in Fig. 1.2
schematically depicts a sensory conflict or the neural
mismatch concept of vertigo and motion sickness.
An active movement leads to stimulation of the sensory organs, whose messages are compared with a
multisensory pattern of expectations calibrated by
earlier experience of motions (central store). The
pattern of expectation is either prepared by the
efference-copy signal, which is emitted parallel to
and simultaneously with the motion impulse, or by
the vestibular excitation during passive transportation in vehicles. If concurrent sensory stimulation
and the pattern of expectation are in agreement,
self-motion is perceived while "space constancy" is
maintained. If, for example, there is no appropriate
visual report of motion, as a result of the field of
view being filled with stationary environmental contrasts (reading in a car), a sensory mismatch occurs
(Dichgans and Brandt 1978; Reason 1978). The
repeated stimulation leads to arearrangement of the
stored pattern of expectation, however, so that a
habituation to the initially challenging stimulation is
obtained within a few days. Sensitisation (or pseudoconditioning) is an increased response to a wide
variety of stimuli. For example, a sensitised animal
responds more vigorously to a mild tactile stimulus
after it has received a painfu1 pinch.

Vestibular compensation and its multiple


mechanisms
The clinical signs, both perceptual and motor, of a
vestibular tone imbalance in the roll plane are ocular
tilt re action, ocular torsion, skew deviation, and tilts
of the perceived visual vertical (Brandt and
Dieterich 1994). A patient with a central vestibular
lesion (Fig. 10.9) presents with a complete ocular tilt
re action to the right, consisting of head tilt, skew
deviation, and ocular torsion. The spontaneous normalisation of ocular torsion, skew deviation, and
apparent tilt of perceived vertical occurs gradually
over several weeks. This time course differs for the
normalisation of ocular motor and perceptual
phenomena (Dieterich and Brandt 1993).
Peripheral vestibular lesions, such as an acute unilateral loss of labyrinthine function as occurs in

57

Management of the dizzy patient

vestibular neuritis (p. 67), cause a distressing tone


imbalance with a spontaneous horizontal rotatory
nystagmus directed away from the affected ear and
an apparent tilt and rotation of the head and body in
the same direction, resulting in a vestibular fall
toward the affected ear, in accordance with the
vestibulospinal reflexes released to counteract the
apparent tilt (Fig. 4.1; Brandt and Daroff 1980; Black
et al. 1989).
Campensation or functional normalisation means
a counterbalancing of any defect of structure or
function. Central compensation of a unilateral
peripheral vestibular loss is considered the prototype of brain plasticity. Postural normalisation in
frogs after a complete unilateral labyrinthectomy
occurs within about 60 days (Flohr et al. 1981) (Fig.
3.4). In the acute stage following haemilabyrinthectomy in the cat, vestibular type 1 neurons show no
spontaneous activity and do not respond to angular
acceleration; some weeks later they recover both

abilities but at higher thresholds (Precht et al. 1966).


The results of occipital lobectomy in rhesus
monkeys demonstrate that visual experience after
labyrinthectomy is essential for the recovery ofVOR
gain, but not for the cessation of spontaneous nystagmus (Fetter et al. 1988). The onset of known
changes in synaptic efficacy of the commissural
vestibular projections on the operated side
(Dieringer and Precht 1977, 1979) are delayed by
about 30 days (see COM in Fig. 3.4d; Kunkel and
Dieringer 1994). Over this period of time, however,
postural normalisation improves already by 50%.
Therefore, commissural changes cannot account for
the early period of postural normalisation. In parallel,
the synaptic efficacy of dorsal root evoked ventral
root responses in the brachial spinal cord of the frog
increases on the operated side as weH (Straka and
Dieringer 1995; Dieringer 1995). These changes,
measured in the isolated spinal cord, parallel more
closely the time course of postural recovery

chronic

acute

40
Ci
(l)
~
c
.Q

100
30
~
~

ro
'> 20

(l)
0)

(l)

"0
"0
eil
(l)

postural

50

eil

..c

10

0
c

15

30

45

60

Time after operation (days)

15

30

45

60
d

Fig.3.4. Time course of postural normalisation after unilateral vestibular lesion in frogs compared with the time course of neural
changes in the brainstem and in the spinal cord on the operated side ofthe same species.a,b Most ofthe postural symptoms presenting acutely after the removal of the labyrinthine organs on the right side a disappear over aperiod of 2 months b. C Time course of
postural normalisation (n = 131) as reported by Flohr et al. (1981). d The curve shown in (c) is expressed in terms of postural recovery
and compared with the time course of an increase in the synaptic efficacy of the commissural vestibular input ((OM; Kunkel and
Dieringer 1994) and of the dorsal root evoked ventral root responses in the isolated brachial spinal root (DR-VRP; Straka and Dieringer
1995). Note that the onset of commissural vestibular changes is delayed and that about 50% of the postural recovery is accomplished
within the first 2 weeks after the lesion. (From Brandt et al. 1997.)

Vertigo

58

(Fig. 3.4d). A unilateral section of the utricular nerve


branch is a necessary and sufficient trigger for both
postural deficits and spinal plastic changes.
Therefore, recovery from vestibular lesions is
neither a simple nor a single process; multiple
processes are involved. Analysis of the mechanisms
of recovery requires a careful comparison of normalisation between parallel phenomena at the
behaviourallevel, on the one hand, and the neuronal
level, on the other. Incongruencies in the time course
and the magnitude of the changes in behaviour and
neuronal activity dearly indicate that multiple
processes of compensation occur in distributed
neuronal networks at different locations and at different times (Straka and Dieringer 1995; Dieringer
1995; Curthoys and Halmagyi 1994).
Various findings in animal experiments have led
to several hypotheses about vestibular compensation
(summarised in Fig. 3.5). Two of these are of special
interest: cerebellar shut-down and increased spinal
input. According to the cerebellar shut-down hypothesis, originally proposed by McCabe and Ryu (1969),
the cerebellum reduces the activity in both vestibular nudei by inhibitory input, thus rebalancing the
activity between the two vestibular nudei. The
reduction of the VOR gain for both directions of
rotation immediately and also 1 year after unilateral
deafferentation (Fetter and Zee 1988) agrees with
this hypothesis. There is also experimental evidence
that increased intrinsic excitability of the ipsilateral
medial vestibular nudei cells is responsible for early
restoration of the resting discharge in these cells and
their subsequent recovery of static vestibular function (Cameron and Dutia 1997).
The increased spinal input hypothesis ascribes
significant static postural changes to the disrupted
activity of the strong descending vestibular spinal
inputs after unilateral vestibular loss. Several studies
have shown that there is a change in the weighting of
spinal afferent input to the vestibular nudei during
compensation. Dieringer et al. (1984) found anatomical evidence for increased spinal afferent projections to the vestibular nudei in the frog following
unilateral vestibular loss. Behavioural studies have
also shown the importance of increased spinal afferent input: cutting cervical dorsal roots causes
decompensation of earlier compensation in squirrel
monkeys (Igarashi et al. 1969). Further, Dichgans et
al. (1973) have reported that the cervico-ocular
reflex is potentiated after bilateral vestibular lesions.

Transmitters of the vestibulo-ocular reflex and


drug-modulated compensation
The relevant anatomical structures of the horizontal
VOR are shown in the scheme in Fig. 3.6 (Raymond

Neural adaptation, contralateral VN


Commissural inhibition
"Cerebellar shutdown "
Gaze changes
Spinal input
Denervation sensitivity
synaptogenesis

24 h

"

2d

axona~ sprouting

7d

2w 4w

h"

2 m 12 m

i
time

Fig. 3.5. Possible mechanisms of restoration of balance in


neural resting activity between the two vestibular nuclei (VN)
after unilateral deafferentation. (From Brandt et al. 1997.)

et al. 1988; Gallagher et al. 1992; Carpenter and


Horik 1992; Phelan et al. 1990; Leigh and Brandt
1993; de Waele et al. 1995). This overview of neurotransmitters and receptor sites involved in the VOR
emphasises the possibly central role that medial
vestibular nu deus neurons play in plasticity of the
vestibular system. As illustrated in the lower right
part of the figure, the medial vestibular nu deus
receives excitatory inputs from the peripheral
vestibular sensory apparatus and inhibitory inputs
from the cerebellum and commissural fibres.
Neurons in the medial vestibular nudeus possess at
least ten different receptors, but the excitatory amino
acid glutamate is the major transmitter used by the
vestibular nerve. Afferent transmission from the hair
cells to the vestibular nerve is regulated by a number
of subtypes of excitatory amino acid receptors as
weIl as by cholinergic and GABAergic efferent feedback from the CNS (Smith and Darlington 1996).
The vestibular nerve and the hair cells engage in
complex transmitter interactions, possibly using
peptides as co-transmitters in order to modify the
sensitivity of post-synaptic receptors (Smith and
Darlington 1996). Under normal conditions there is
a balance between the release of excitatory amino
acids to medial vestibular nudeus neurons and the
opposing inhibitory transmitters. Under pathological conditions such as unilateral vestibular loss,
inhibitory influences may dominate, causing an
imbalance, which results in dinical signs of rotatory
vertigo, nystagmus, and postural imbalance.
Pharmacological and metabolic studies suggest
that the process of central compensation for peripheral vestibular lesions is both dynamic and fragile:
alcohol, phenobarbital, chlorpromazine, diazepam,
and ACTH antagonists retard compensation, whereas
caffeine, amphetamines, steroids, and ACTH accelerate

59

Management of the dizzy patient

eye museIes

I lateral redu.

oeulomotor nueleus

abdueens nueleus

Flocculus.
Nel. fa stigii

n1 I

-g>

"6

commissural fibres

vestibular nuelei

vestibular afferent fibre

-----_.,

GABAA
Receptor
~l and
Opioid Reeeptor
NAdr

~~:~::ar
I
I

IG
-::--'lu--- ta-m-a-te'l

NMDA and

~::p~~A-

fibre

'_ .. - - .... - - - - - - .. - _................................................................... ~

Fig.3.6. Summary of excitatory and inhibitory neurotransmitters as weil as receptor sites of the vestibulo-ocular reflex. This figure
emphasises that there are many receptor sites, especially within the (medial) vestibular nucleus. ACh: acetylcholine, DA: dopamine
receptor, GABA: gamma-aminobutyric acid, H2 : H2 histamine receptor, N: nicotinic receptor, NMDA: N-methyl-D-aspartate, M: muscarinic receptor, NE: norepinephrine receptor, 5-HT 3 : serotonin receptor. (Modified after Raymond et al. 1988; Gallagher et al. 1992;
Carpenter and Horik 1992; Phelan et al. 1990; de Waele et al. 1995.)

60

it (Zee 1988; Darlington et al. 1990; Gilchrist et al.


1990; Darlington and Smith 1992; Yamanaka et al.
1995). However, there are also reports available that
even high doses of diazepam before and following
unilaterallabyrinthectomy do not result in impaired
compensation of spontaneous nystagmus in guineapigs (Martin et al. 1996). Furthermore, gangliosides
(Petrosini 1987), thyrotropine-releasing hormones
(Ishii and Igarashi 1986), betahistine (Tighilet et al.
1995), and extract from Gingko biloba (Ez-Zaher and
Lacour 1988) were reported to promote vestibular
compensation in humans and animals (Smith and
Darlington 1994; Hamann 1988). The efficacy of
these agents in humans is probable for melanotropic
peptides and Gingko biloba extract, but still has to be
proven in controlled studies (Smith and Darlington
1994). We share the observation reported by Thmke
et al. (1990) that alcohol intake may transiently
decompensate central compensation of a peripheral
vestibular lesion. This is also supported by animal
experiments: after completely recovering from postural imbalance, unilaterally labyrinthectomised
guinea-pigs suffer a relapse after ingesting alcohol
(Schaefer et al. 1978).

Substitution of vestibular function


Vestibular compensation is less perfect than generally
believed. For instance, after acute unilateral vestibular deafferentation, which occurs in vestibular neuritis, the process of normalisation is impressive for
static conditions in the absence of head motion: the
initial rotatory vertigo, spontaneous nystagmus, and
postural imbalance subside. Compensation is, however, less impressive for dynamic conditions,
especially when the vestibular system is exposed to
high-frequency head accelerations (Curthoys and
Halmagyi 1994; Halmagyi et al. 1990). The dynamic
disequilibrium, i.e. VOR asymmetry, causes oscillopsia, the illusory movement of the environment due to
excessive slip of images upon the retina during fast
head movements or walking, because after uni- and
bilateral peripheral vestibular lesions the VOR
cannot generate fast compensatory eye rotations
during high-frequency head rotations. The dynamic
vestibular tone imbalance can be detected clinically
by provoking a directional head-shaking nystagmus
(Hain et al. 1987) or by bedside testing of the VOR
with rapid head rotation (Halmagyi and Curthoys
1988).
The vestibular system is considered a good exampIe of neural plasticity, since the VOR gain changes
with altered visual input. Despite the powerful adaptive control of the VOR gain, however, there are only

Vertigo
comparatively small changes of dynamic vestibular
function following unilateral or bilateral vestibular
loss. How can this paradox be explained? The direct
elementary three-neuron VOR pathway - essential
for the short latency properties 16 ms) of the VOR
- can hardly be modified (Lisberger and Pavelko
1988; Snyder et al. 1992). In contrast, the parallel network to the indirect oligosynaptic VOR pathway
(Fig. 3.7) is capable of gain changes via the feed-forward or open-loop control system (Lisberger et al.
1984; Lisberger and Sejnowski 1992). Thus, the
asymmetrical responses of individual semicircular
canals (described by Ewald's second law), which are
usually concealed by the bilateral interaction
between the two labyrinths, cause after unilateral
deafferentation persistent asymmetries of the
dynamic VOR gain [< 0.5 on the affected side
(Curthoys and Halmagyi 1994; Halmagyi et al. 1990)].
It then becomes clear that a further mechanism
must subserve the functionally insufficient compensation: this mechanism is substitution (Curthoys
and Halmagyi 1994). In case of a deficient VOR, a
refixation saccade ("catch-up saccade") substitutes
for the lack of compensatory slow movement even
in frogs (Dieringer 1988). Further, patients learn
new behavioural strategies (restrietion of the head
movements toward the affected ear, making isolated
eye instead of combined eye-head movements) or
alter the relative weights of inputs to the gaze and
posture control systems (Angelaki et al. 1993). The
same is true for multisensory interaction: vision
and proprioception may substitute parts of the
missing vestibular input for postural control. An
asymmetrie increase in cervical muscle spindie
input (restricted to the affected side) following
vestibular neuritis has been demonstrated by the
differential effects of neck muscle vibration (Strupp
et al. 1998a). This increase gradually builds up over
weeks.
Thus, the so-called simple and complete vestibular compensation for peripheral deficits is only a legend. Nevertheless, the vestibular system provides an
excellent and attractive model for investigations of
neural and behavioural plasticity in humans and
animals. Certain distinct features have advantages
(Brandt et al. 1997):
1. the periperal vestibular lesion can be precisely
located, is restricted, and easy to reproduce without disturbing central parts of the vestibular system, which are important for plasticity;
2. the recovery of function - as weIl as its time
course - can be measured quantitatively at different levels (vestibulospinal reflex, vestibuloocular reflex, and perception);

Management of the dizzy patient

FLOCCULUS

61

Visuallnput

Oculomotor Input

Oculomotor
Nuclei

FEEDBACK

EYE MOVEMENTS
Cell

Flocculus - Target Neuron

Vestibular Input

Second order

First or second order

First order

VESTIBULAR INPUT
Fig.3.7. Hypothetical model of vestibulo-ocular adaptation. The flocculus receives information from the vestibular, visual, and oculomotor systems. These signals may be used by the flocculus both to compute errors in the vestibulo-ocular reflex and via flocculustarget neurons to change the gain and phase of the VOR. (Adapted from Lisberger et al. 1984.)

3. the anatomy, physiology, and functions of the


VOR network have been intensively studied.

References

Vestibular compensation is usually considered a


central "repair mechanism" for a vestibular tone
imbalance secondary to a peripheral vestibular loss.
However, central compensation is also possible for
central vestibular tone imbalances. This is best
demonstrated by the cessation of nystagmus and lateropulsion in Wallenberg's syndrome (p. 309). It is
still poorly understood which central vestibular syndromes can be compensated and which cannot.
Upbeat and downbeat nystagmus may serve as
an example. Acquired upbeat nystagmus is rarely
permanent, whereas acquired downbeat nystagmus
may be permanent (p. 199).
A lot is now known about vestibular compensation. When it is fully understood, we shall be able to
understand most mechanisms underlying plasticity
of the central nervous system.

Angelaki OE, Bush GA, Perachia AA (1993) Two-dimensional


spatiotemporal coding of linear acceleration in vestibular
nuclei neurons. J Neurosci 13:1403-1417
Averbuch-Heller L, Tusa RJ, Fuhry L, Rottach KG, Ganser GL,
Heide W, Bttner U, Leigh RJ (1997) A double-blind controlled
study of gabapentin and baclofen as treatment for acquired
nystagmus. Ann NeuroI41:818-825
Bechterew W von (1883) Ergebnisse der Ourchschneidung des N.
acusticus, nebst Errterung der Bedeutung der semicirculren
Kanle fr das Krpergleichgewicht. Pflgers Arch 30:312-347
Benecke JE (1994) Surgery for non-Meniere's vertigo. Acta
Otolaryngol (Stockh) SuppI513:37-39
Bergenius J, dkist LM (1996) Transtympanic aminoglycoside
treatment of Meniere's disease. In: Baloh RM, Halmagyi GM
(eds) Disorders of the vestibular system. Oxford University
Press, New York, Oxford, pp 575-582
Black FO, Shupert CL, Peterka RJ, Nashner LM (1989) Effects of
unilateralloss of vestibular function on the vestibulo-ocular
reflex and postural contro!. Ann Otol Rhinol Laryngol
98:884-889
Bhmer A, Fisch U (1993) Bilateral vestibular neurectomy for
treatment ofvertigo. Otolaryngol Head Neck Surg 109:101-107

62
Brackmann DE (1996) Surgical procedures: endolymphatic shunt,
vestibular nerve section and labyrinthectomy. In: Baloh RW,
Halmagyi GM (eds) Disorders of the vestibular system. Oxford
University Press, Oxford, PP 551-562
Brandt T (1993) Management of acute peripheral vestibular disorders. Eur Neurol 33:337-340
Brandt Th, Paulus W (1989) Postural retraining in exceptional
populations. In: Wollacott M, Shumway-Cook A (eds) Posture
and gait across the lifespan. University of South Carolina Press,
pp 299-319
Brandt T, Daroff RB (1980) The multisensory physiological and
pathological vertigo syndromes.Ann NeuroI7:195-203
Brandt T, Dieterich M (1994) Vestibular syndromes in the roll
plane: topographic diagnosis from brainstem to cortex. Ann
NeuroI36:337-347
Brandt Th, Dichgans J, Wagner W (1974) Drug effectiveness on
experimental optokinetic and vestibular motion sickness.
Aerospace Med 45:1291-1297
Brandt Th, Krafczyk S, Malsbenden I (1981) Postural imbalance
with head extension: improvement by training as a model for
ataxia therapy.Ann NY Acad Sei 374:646-649
Brandt Th, Bchele W, Krafczyk S (1986) Training effects on
experimental postural instability: a model for clinical ataxia
therapy. In: Bles W, Brandt Th. (eds) Disorders in posture and
gait. Elsevier, Amsterdam, pp 353-365
Brandt T, Strupp M, Arbusow V, Dieringer N (1997) Plasticity of
the vestibular system: central compensation and sensory substitution for vestibular deficits. Ann Neurol 73:297-309
Bchele W, Knaup H, Brandt Th (1984) Time course of training
effects on balaneing on one foot. Acta Otolaryngol (Stockh)
SuppI406:140-142
Cameron SA, Dutia MB (1997) Cellular basis of vestibular
compensation: changes in intrinsic excitability of MVN neurones. NeuroReport 8:2595-2599
Carpenter DO, Horik N (1992). Neurotransmitter and peptide
receptors on medial vestibular nucleus neurons. Ann NY Acad
Sei 656:668-686
Courjon JH, Jeannerod M (1979) Visual substitution of
labyrinthine defect. In: Granit R, Pompeiano 0 (eds) Progress
in brain research, vol 50. Reflex control of posture and locomotion. Elsevier, Amsterdam, pp 783-792
Courjon JH, Jeannerod M, Ossuzio I, Schmid R (1977) The role of
vision in compensation of vestibulo-ocular reflex after hemilabyrinthectomy in the cat. Exp Brain Res 28:235-248
Curthoys IS, Halmagyi GM (1994) Vestibular compensation: a
review of the oculomotor, neural, and clinical consequences of
unilateral vestibular loss. J Vestib Res 5:67-107
Darlington CL, Smith PF (1992) Pre-treatment with a Ca 2+ channel antagonist facilitates vestibular compensation. NeuroReport
3:143-145
Darlington CL, Smith PF, Hubbard JI (1990) Guinea pig medial
vestibular nucleus neurons in vitro respond to ACTH (4-10) at
picomolar concentrations. Exp Brain Res 82:637-640
de Waele C, Mhlethaler M, Vidal PP (1995) Neurochemistry of
the central vestibular pathways. Brain Res Rev 20:24-46
Dichgans J, Brandt T (1978) Visual-vestibular interaction: effects
on self-motion perception and postural contro!. In: Held R,
Leibowitz HW, Teuber HL (eds) Handbook of sensory physiology, vo!. 8. Perception. Springer, New York, pp 755-804
Dichgans J, Bizzi E, Morasso P, Tagliasco V (1973) Mechanisms
underlying recovery of eye-head co ordination following bilaterallabyrinthectomy in monkeys. Exp Brain Res 18:548-562
Dieringer N (1988) Immediate saccadic substitution for defieits in
dynamic vestibular reflexes of frogs with selective peripheral
lesions. Prog Brain Res 76:403-409
Dieringer N (1995) "Vestibular compensation": neural plasticity
and its relations to functional recovery after labyrinthine

Vertigo
lesions in frogs and other vertebrates. Prog Neurobiol
46:97-129
Dieringer N, Precht W (1977) Modified synaptic input in chronically deafferented neurons. Nature 269:431-433
Dieringer N, Precht W (1979) Mechanisms of compensation for
vestibular defieits in the frog. I. Modification of the exeitatory
commissural system. Exp Brain Res 36:311-328
Dieringer N, Knzle H, Precht W (1984) Increased projection of
ascending dorsal root fibers to vestibular nuclei after hemilabyrinthectomy in the frog. Exp Brain Res 55:574-578
Dieterich M, Brandt T (1993) Thalamic infarctions: differential
effects on vestibular function in the roll plane (35 patients).
Neurology 43:1732-1740
Dieterich M, Straube A, Brandt T, Paulus W, Bttner U (1991) The
effects of baclofen and anticholinergic drugs on upbeat and
downbeat nystagmus. J Neurol Neurosurg Psychiatry
54:627-632
Dorland's Illustrated Medical Dictionary (1994) Saunders,
Philadelphia
Ehrenberger K, Felix D (1996) Receptor pharmacological models
for the therapy of labyrinthine vertigo. Acta Otolaryngol
(Stockh) 116:189-191
Ewald JR (1892) Physiologische Untersuchungen ber das
Endorgan des N. octavus. Bergmann, Wiesbaden
Ez-Zaher L, Lacour M (1988) Effects of an extract of Ginkgo biloba
on vestibular compensation in the cat. In: Claussen CF, Kirtane
MV, Schlitter K (eds) Vertigo, nausea, tinnitus and hypoacusia
in metabolic dis orders. Elsevier, Amsterdam, pp 595-600
Fetter M, Zee DS (1988) Recovery from unilaterallabyrinthectomy
in rhesus monkeys. J Neurophysiol 59:370-393
Fetter M, Zee DS, Proctor LR (1988) Effect of lack of vision and of
oceipitallobectomy upon recovery from unilateral labyrinthectomy in rhesus monkey. J Neurophysiol 59:349-407
Flohr H, Bienhold H, Abeln W, Macskovics I (1981) Concepts of
vestibular compensation. In: Flohr H, Precht W (eds) Lesioninduced neuronal plastieity in sensorimotor systems. Springer,
NewYork,pp 153-172
Foster C, Baloh RW (1996) Drug therapy for vertigo. In: Baloh RW,
Halmagyi GM (eds) Disorders of the vestibular system. Oxford
University Press, Oxford, pp 541-550
Gacek RR, Gacek MR (1996) Comparison of labyrinthectomy and
vestibular neurectomy in the control of vertigo. Laryngoscope
106:225-230
Gallagher JP, Phelan KD, Shinnick-Gallagher P (1992) Modulation
of exeitatory transmission at the rat medial vestibular nucleus
synapse. Ann NY Acad Sei 656:630-644
Gilchrist DP, Smith PF, Darlington CL (1990) ACTH(4-1O) accelerates ocular motor recovery in the guinea pig following vestibular deafferentation. Neurosei Lett 118:14-16
Gonshor A, Melvill Jones G (1976) Extreme vestibulo-ocular adaptation induced by prolonged optical reversal of vision. J Physiol
(Lond) 256:381-414
Hain TC, Fetter M, Zee DS (1987) Head-shaking nystagmus in
patients with unilateral peripheral vestibular lesions. Am J
OtolaryngoI8:36-47
Haines RF (1974) Effect ofbed rest and exereise on body balance.
J Appl Physiol 36:323-327
Halmagyi GM (1994) Vestibular insufficiency following unilateral
vestibular deafferentation. Aust J Otolaryngol 1:510 -512
Halmagyi GM, Curthoys IS (1988) A clinical sign of canal paresis.
Arch NeuroI45:737-739
Halmagyi GM, Rudge P, Gresty MA (1980) Treatment of periodic
alternating nystagmus. Ann NeuroI8:609-611
Halmagyi GM, Curthoys IS, Cremer PD, Henderson CJ, Todd MJ,
Staples MJ, D'Cruz DM (1990) The human horizontal vestibuloocular reflex in response to high -acceleration stimulation
before and after unilateral vestibular neurectomy. Exp Brain
Res 81:479-490

Management of the dizzy patient


Halmagyi GM, Curthoys IS, Todd MJ, D'Cruz DM, Cremer PD,
Henderson q, Staples MJ (1991) Unilateral vestibular neurectomy in man causes a severe permanent horizontal vestibulo-ocular reflex deficit in response to high-acceleration ampullofugal
stimulation. Acta Otolaryngol (Stockh) SuppI481:411-414
Hamann KF (1988) Rehabilitation of patients with vestibular disorders. HNO 36:305-307
Herdman SJ (1996) Vestibular rehabilitation. In: Baloh RW,
Halmagyi GM teds) Disorders of the vestibular system. Oxford
University Press, Oxford, pp 583-597
Herdman SJ, Clendaniel RA, Mattox DE, Holiday M, Niparko JK
(1995) Vestibular adaptation exercises and recovery: acute stage
following acoustic neuroma resection. Otolaryngol Head Neck
Surg 113:71-77
Horak FB, Jones-Rycewicz C, Black 0, Shumway-Cook A (1992)
Effects of vestibular rehabilitation on dizziness and imbalance.
Otolaryngol Head Neck Surg 106:175-180
Igarashi M (1986) Compensation for peripheral vestibular disturbances - animal studies. In: Bles W, Brandt Th teds) Disorders
of posture and gait. Elsevier, Amsterdam, pp 337-351
Igarashi M, Alford BR, Watanabe T, Maxian PM (1969) Role of
neck proprioception for the maintenance of dynamic bodily
equilibrium in the squirrel monkey. Laryngoscope
79:1713-1727
Igarashi M, Ishikawa K, Ishii M, Yamane H (1988) Physical exercise
and balance compensation after total ablation of vestibular
organs. Prog Brain Res 76:395-401
Ishii M, Igarashi M (1986) Effect of thyrotropin-releasing hormone on vestibular compensation in primates. Am J
OtolaryngoI7:177-180
Kanayama R, Bronstein AM, Gresty MA, Brookes GB, Faldon ME,
Nakamura T (1995) Perceptual studies in patients with vestibular neurectomy. Acta Otolaryngol (Stockh) Suppl 520:408-411
Kemink JL, Telian SA, El-Kashlan H, Langman AW (1991)
Retrolabyrinthine vestibular nerve section: efficacy in disorders
other than Meniere's disease. Laryngoscope 101:523-528
Kingma H, Bonink M, Meulenbroeks A, Konijnenberg H (1997)
Dose-dependent effect of betahistine on the vestibulo-ocular
reflex: a double-blind placebo controlled study in patients with
paroxysmal vertigo.Acta Otolaryngol (Stockh) 117:641-646
Krebs DE, Gill-Body KM, Riley PO, Parker SW (1993) Doubleblind placebo controlled trial of rehabilitation for bilateral
vestibular hypofunction: preliminary report. Otolaryngol Head
Neck Surg 109:735-741
Kunkel AW, Dieringer N (1994) Morphological and electrophysiological consequences of unilateral pre- versus post -ganglionic
vestibular lesions in the frog. J Comp Physiol (A) 174:621-632
Lacour M (1984) Reapprentissage et periode postoperatoire sensible dans la restauration des fonctions nerveuses. Exemple de la
compensation vestibulaire et implications cliniques. Ann
Oto-Laryngol (Paris) 101:177-187
Lacour M, Xerri C (1981) Vestibular compensation: new perspectives. In: Flohr H, Precht W (eds) Lesion-induced neuronal plasticity in sensorimotor systems. Springer, Berlin, pp 240-253
Lacour M, Roll JR, Apaix M (1976) Modifications and development of spinal reflexes in the alert baboon (Papio papio) following unilateral vestibular neurectomy. Brain Rev 113:255-269
Leigh RJ, Brandt T (1993) Areevaluation of the vestibulo-ocular
reflex: new ideas of its purpose, properties, neural substrate,
and dis orders. Neurology 43:1288-1295.
Leigh RJ, Averbuch -Heller L, Tomsak RL, Remler BF, Yaniglos SS,

63
Dell'Osso LF (1994) Treatment of abnormal eye movements
that impair vision: strategies based on current concepts of
physiology and pharmacology. Ann NeuroI36:129-141
Lisberger SG, Pavelko TA (1988) Brain stern neurons in modified
pathways for motor learning in the primate vestibulo-ocular
reflex. Science 242:771-773
Lisberger SG, Sejnowski TJ (1992) Motor learning in a recurrent
network model based on the vestibulo-ocular reflex. Nature
360:159-161
Lisberger SG, Miles FA, Zee DS (1984) Signals used to compute
errors in monkey vestibulo-ocular reflex: possible role of flocculus. J NeurophysioI52:1140-1153
Manning C, Scandale L, Manning EI, Gengo FM (1992) Central
nervous system effects of meclicine and dimenhydrinate: evidence of acute tolerance to antihistamines. J Clin Pharmacol
32:996-1002
Martin J, Gilchrist DPD, Smith PF, Darlington CL (1996) Early
diazepam treatment following unilaterallabyrinthectomy does
not impair vestibular compensation of spontaneous nystagmus
in the guinea pig. J Vestib Res 6: 135-139
McCabe BF, Ryu JH (1969) Experiments on vestibular compensation. Laryngoscope 79: 1728-1736
Nomura Y, Okuno T, Mizuno M (1993) Treatment of vertigo using
laser labyrinthectomy. Acta Otolaryngol (Stockh) 113:261-262
Nomura Y, Ooki S, Kukita N, Young Y-H (1995) Laser labyrinthectomy. Acta Otolaryngol (Stockh) 115:158-161
Petrosini L (1987) Behavioural recovery from unilateral vestibular
lesion is facilitated by GM, ganglioside treatment. Behav Brain
Res 23:117-126
Phelan KD, Nakamura J, Gallagher JP (1990). Histamine depolarizes rat medial vestibular nucleus neurons recorded intracellularly in vitro. Neurosci Lett 109:287-292
Pohl DV (1996) Surgical procedures for benign positional vertigo.
In: Baloh RW, Halmagyi GM (eds) Disorders of the vestibular
system. Oxford University Press, Oxford, pp 563-582
Precht W, Shimazu H, Markharn CH (1966) A mechanism of central compensation of vestibular function following hemilabyrinthectomy. J Neurophysiol 29:996
Raymond J, Dememes D, Nieoullon A (1988) Neurotransmitters in
vestibular pathways. Prog Brain Res 76:29-43
Reason JT (1978) Motion sickness adaptation: a neural mismatch
model. J R Soc Med 71:819-829.
Schaefer KP, Wilhelms G, Meyer DL (1978) Der Einflu von
Alkohol auf die zentralnervsen Ausgleichvorgnge nach
Labyrinthausschaltung. Z Rechtsmed 81:249-260
Smith PF, Darlington CL (1994) Can vestibular compensation be
enhanced by drug treatment? J Vestib Res 4:169-179
Smith PF, Darlington CL (1996) Recent advances in the pharmacology of the vestibulo-ocular reflex system. TINS 17:421-427
Snyder LH, Lawrence DM, King WM (1992) Changes in vestibuloocular reflex (VOR) anticipate changes in vergence angle in
monkey. Vision Res 32:569-575
Starck M, Albrecht H, Pllman W, Straube A, Dieterich M (1997)
Drug therapy for acquired pendular nystagmus in multiple
sclerosis. J NeuroI244:9-16
Straka H, Dieringer N (1995) Spinal plasticity after hemilabyrinthectomy and its relation to postural recovery in the
frog. J Neurophysiol 73:1617-1631
Strupp M, Arbusow V, Dieterich M, Sautier W, Brandt T (l998a)
Perceptual and oculomotor effects of neck muscle vibration in

64
vestibular neuritis. Ipsilateral somatosensory substitution of
vestibular function. Brain 121:677-685
Strupp M, Arbusow V, Maag KP, GaU C, Brandt T (1998b)
Vestibular exercises improve central vestibulo-spinal compensation after vestibular neuritis. Neurology 51:838-844
Szturm T, Ireland DJ, Lessing-Turner M (1994) Comparison of different exercise programs in the rehabilitation of patients with
chronic peripheral vestibular dysfunction. J Vestib Res
6:461-479
Taylor HL, Henschel A, Brozek J, Keys A (1949) Effect ofbed rest
on cardiovascular function in work performance. J Appl Physiol
2:223-239
Thmke F, Vogt T, Hopf HC (1990) Alcohol-dependent unilateral
vestibular impairment persisting after a closed head injury. J
NeuroI237:326-327
Thompson RF, Spencer WA (1966) Habituation: a model phenomenon for the study of neuronal substrate behaviour. Psychol Rev
73:16-43.
Thomsen J, Bretlau P, Tos M, Johnsen NJ (1981) Placebo effect in
surgery for Meniere's disease. Arch Otolaryngol107:271-277
Tighilet B, Leonard J, Lacour M (1995) Betahistine dihydro-

Vertigo
chloride treatment facilitates vestibular compensation in the
cat. J Vestib Res 5:53-66
Timmerman H (1994) Pharmacotherapy ofvertigo: any news to
be expected? Acta Otolaryngol (Stockh) SuppI513:28-32
Vaisrub N (1981) Summary statement to "Placebo effect for
Meniere's disease sac shunt surgery disputed". (letter to the editor) Arch Otolaryngoll07:773-774
Waespe W, Cohen B, Raphan T (1985) Dynamic modification of
the vestibulo-ocular reflex by the nodulus and uvula. Science
228:199-202
Xerri C, Lacour M (1980) Compensation des deficits posturaux et
cinetiques apres neurectomie vestibulaire unilaterale chez le
chat.Acta Otolaryngol (Stockh) 90:414-424
Yamanaka T, Sasa M, Amano T, Miyahara H, Matsunaga T (1995)
Role of glucocorticoid in vestibular compensation in relation to
activation of vestibular nucleus neurons. Acta Otolaryngol
Suppl (Stockh) 519:168-172
Zee DS (1988) The management of patients with vestibular disorders. In: Barber HO, Sharpe JA (eds) Vestibular disorders. Year
Book, Chicago, pp 254-274

SECTION B
Vestibular nerve and
labyrinthine dis orders

Vestibular neuritis

Acute unilateral (idiopathic) vestibular paralysis,


also known as vestibular neuritis (VN), is the third
most common cause of peripheral vestibular vertigo.
Jt accounts for about 5% of the patients referred to a
neuroLogical dizziness unit. It was first described by
Ruttin in 1909 and later by Nylen (1924). The term
"vestibular neuronitis," coined by Hall pike (1949)
and Dix and Hallpike (1952), shouLd be repLaced
by "vestibular neuritis," because there is strong
evidence that the ganglion cells themselves are not
inflamed, but rat her parts of the nerve, i.e. the
neurite.
The chief symptom is the acute on set of prolonged severe rotatory vertigo, associated with spontaneous horizontal-rotatory nystagmus, postural
imbalance, and nausea without concomitant auditory dysfunction. Caloric testing invariably shows ipsilateral hypo- or non-responsiveness (as a sign of
horizontal semicircular canal paresis). Epidemie
occurrence of the condition, the frequency of preceding upper respiratory tract infections, a smaH
number of post-mortem studies that fou nd ceH
degeneration of one or more vestibular nerve
trunks, and the demonstration of latent herpes simplex virus type I in human vestibular ganglia - all
suggest that the cause may be a viral infection (or
reactivation) of the vestibular nerve, similar to those
producing BeLl's palsy and sudden sensorineural
hearing loss.
V is most Iikely a partial rather than a complete
vestibular paresis, with predominant involvement of
the horizontal and anterior semicircular canals
(sparing the posterior semicircular canal). The condition mainly affects adults, ages 30 to 60, and has a
natural history of gradual recovery within 1 to 6
weeks. Recovery is a product of combined (I)
peripheral restoration of labyrinthine function (frequently incomplete); (2) (vestibular contralateral),
somatosensory, and visual substitution for the unilateral vestibular deficiti and (3) central compensation of the vestibular tone imbalance (aided by
physical exercise).
67

Diagnosis of VN is based on the simple assessment of an acute vestibular tone imbalance associated
with a unilateral peripheral vestibular loss (bedside
testing of high-frequency vestibular ocular reflex;
caloric testing) after clinical exclusion of a central
neurological disorder. As this diagnostic procedure
lacks selectivity, pathological processes other than
V which also cause an acute unilateral loss of
peripheral labyrinthine function may be wrongly
labelled. Thus, the term VN does not describe a welldefined clinical entity, but rather a syndrome in
which peripheral vestibular paralysis can have a
nurnber of possible causes (usuaLly viral or vascular). Some authors have proposed other sites for the
lesion: peripherallabyrinth, vestibular nerve, or the
insertion site oE the root of the eighth nerve into the
ponto-medullary brainstem (here an MS plaque can
mimic VN). Differential diagnosis includes all other
causes oE acute loss of peripherallabyrinthine function (p. 72).

The clinical syndrome


Key signs and symptoms of VN (Fig. 4.I) are an
acute onset of sustained
rotatory vertigo (contraversive) with pathological adjustments of perceived straight-ahead and
subjective vertical (ipsiversive),
2. postural imbalance with Romberg fall and pastpointing (ipsiversive),
3. horizontal-rota tory spontaneous nystagmus
(contraversive) with oscillopsia, and
4. nausea and vomiting.
1.

Ocular motor evaluation reveals apparent horizontal saccadic pursuit, gaze-evoked nystagmus
toward the fast phase of the spontaneous nystagmus,
and a directional preponderance of optokinetic

Vertigo

68

nystagmus (contraversive to the lesion), all of which


are secondary to the spontaneous nystagmus
indicating vestibular tone imbalance in yaw and roll
planes. Hearing loss is not a typical feature of the
condition, and the detection of any neurological
deficit beyond the above indicated signs and symptoms should raise doubts about the diagnosis ofVN.
A suspected diagnosis can be hardened by
demonstrating a unilateral deficit in vestibuloocular reflex bedside testing (p. 39) and, more definitively, hypo- or unresponsiveness in bithermal
caloric testing (Fig. 4.2; horizontal semicircular
canal paresis of the labyrinth opposite to the fastphase of the spontaneous nystagmus). There is, however, no pathognomonic test or sign for VN as a
clinical entity. In a strict sense, only an acute unilateral peripheral vestibular hypofunction with horizontal semicircular canal paresis can be diagnosed
by the proposed procedure. The thus defined group
has many of the clinical features described below in
common and does not require further apparative or
invasive diagnostics, although some patients may
have a different aetiology.

!~
L

~
Vertigo
~

Ocular torsion
Subjective visual vertical
Subjective straight ahead
Fig.4.1. Ocular signs, perception, and posture in the acute
stage of right-sided vestibular neuritis. Spontaneous vestibular
nystagmus is always horizontal-rotatory away from the side of
the lesion (best observed with Frenzel's glasses). The initial perception of apparent body motion (vertigo) is also directed away
from the side of the lesion, whereas measurable ocular torsion
and body destabilisation (Romberg fall) are always toward the
side of the lesion. The latter are the compensatory vestibulo-ocular and vestibulospinal reactions to the apparent tilt. The same is
true for adjustments of perceived vertical and subjective straight
ahead. (From Brandt and Dieterich 1995.)

Fig.4.2. Eye signs in the acute stage of vestibular neuritis right


(arrow) . Spontaneous nystagmus is always horizontal-rotatory
away from side of the lesion. It is best observed with Frenzel's
glasses, since fixation largely suppresses nystagmus (top). With lateral gaze and fixation of a stationary target, spontaneous nystagmus is inhibited when gaze is directed toward the affected ear and
increased when gaze is directed toward the unaffected ear (middie). Thermic irrigation of the external auditory canal (caloric test)
demonstrates unresponsiveness ofthe affected right horizontal
semicircular canal but normal responses of the left horizontal
semicircular canal (bottom). Spontaneous vestibular nystagmus to
the left causes a directional bias of the recorded eye movements
during thermic irrigation.

Vertigo and posture


In VN the fast phase of the spontaneous rotatory
nystagmus (Fig. 4.2) and the initial perception of
apparent body motion are directed away from the
side of the lesion, and the postural reactions initiated by vestibulospinal reflexes are usually opposite to
the direction of vertigo. These result in both the
Romberg fall and in past-pointing toward the side of
the lesion. Patients with this type of vertigo often
make confusing and contradictory statements about
the directionality of their symptoms. In actual fact,
there are two sensations of opposite directions (see
Vestibular falls; p. 13), and the patient may be

69

Vestibular neuritis

describing one or the other. The first is a purely


subjective sense of self-motion in the direction of
the nystagmus fast phases, which is not associated
with any measurable body sway. The second is the
compensatory vestibulospinal re action resulting in
objective, measurable destabilisation and a possible
Romberg fall in the direction opposite to the fast
phases (Brandt and Daroff 1980). Subjective
straight-ahead and tilts of perceived vertical can be
determined psychophysically as the perceptual consequence of vestibular tone imbalance in yaw (horizontal semicircular canal) and roll (utricle and
anterior semicircular canal). The direction of pathological deviation - as adjusted by the patient - and
the Romberg fall are ipsiversive to the lesioned ear.
The severity of tone imbalance can be measured in
degrees; it is thus possible to assess quantitatively
the recovery of spatial disorientation during the
course of the illness. Furthermore, the efficacy of
physical and medical treatment of the condition can
also be measured in this way (Strupp et al. 1998b).

Eye movements
The nystagmus is always rotatory-horizontal (beating clockwise-Ieft or counterclockwise-right); a
purely linear nystagmus is not compatible with the
diagnosis. The nystagmus is typically reduced in
amplitude by fixation (fixation suppression) and
enhanced by eye closure or Frenzel's (high plus)
lenses. According to Alexander's law, amplitude and
slow-phase velocity are increased with gaze shifts
toward the fast phase, and decreased with gaze shifts
toward the slow phase of the nystagmus. This may
mimic unilateral gaze-evoked nystagmus in a patient
with moderate, spontaneous nystagmus that is completely suppressed by fixation straight ahead but still
present with the gaze directed toward the fast phase.
Using a motor-driven 3D rotating chair, Fetter and
Dichgans (1996) studied 3D properties of the
vestibulo-ocular reflex in 16 patients in the acute
stage ofVN. Their measurements support the view
that VN is a partial rather than a complete unilateral
vestibular lesion (Bchele and Brandt 1988) and that
this partiallesion affects the superior division of the
vestibular nerve including the afferents from the
horizontal and anterior semicircular canals (Fetter
and Dichgans 1996): "In all patients, spontaneous
nystagmus axes clustered between the direction
expected with involvement of just one horizontal
semicircular canal and the direction expected with
combined involvement of the horizontal and anterior
semicircular canals on one side. Likewise, dynamic
asymmetries were found only during rotations about
axes which stimulated the ipsilesional horizontal

or ipsilesional anterior semicircular canals. No


asymmetry was found when the ipsilesional posterior
semicircular canal was stimulated:' This analysis was
based on physiological data that electrical stimulation of single semicircular canal nerves elicits eye
movements in the plane of the canal (Suzuki and
Cohen 1964) and that combinations of canallesions
should result in a direction of spontaneous nystagmus, which reflects the weighted vector sum of the
axes of the involved canals.
A significant directional preponderance of optokinetic nystagmus (OKN) may be another consequence of the peripherallesion (Ohm 1932; Jung
and Mittermaier 1939), and not a result of involvement of the brainstem or cerebellum. These vestibularIy induced differences in OKN-slow phase velo city
can be as large as 70%, and are due to enhancement
toward the side of the lesion and depression in the
opposite horizontal direction (Brandt et al. 1978).
The interaction is not purely additive or subtractive;
a feedforward optokinetic gain control of the
vestibular component (multiplication) is involved
before the two signals are combined (Fig. 4.3).
Skew deviation (ipsilesional eye undermost) and
ocular torsion with vertical and oblique diplopia
have also been described in some patients with acute
VN (Safran et al. 1994; Vibert et al. 1996) as weIl as
perceived tilts of visual vertical (Bhmer and
Rickenmann 1995). This indicates a vestibular tone
imbalance in the roll plane (see Central vestibular
dis orders in roll plane; p. 175) induced by involveme nt of the anterior semicircular canal, otolith function, or both. The superior division of the vestibular
nerve innervates not only the cristae of the horizontal and anterior semicircular canals but also the
maculae of the utricle and the anterosuperior part of
the saccule.It is possible that alesion of only the
superior division results in ocular torsion and tilts of
the visual vertical, whereas alesion of both the
superior and the inferior divisions of the eighth
nerve results in a complete ocular tilt re action and
head tilt. We have seen evidence for the latter in a
patient with herpes zoster oticus, which manifested
with an ocular tilt reaction and showed a contrast
enhancement of the superior and inferior parts of
the eighth nerve on MRI (Arbusow et al. 1998).

Caloric testing
The principal diagnostic marker of the disease is an
initial paresis of the horizontal semicircular canal on
the affected side; this can be demonstrated by caloric
tests (Fig. 4.2). Meran and Pfaltz (1975) reported that
2 weeks after onset of vestibular neuritis, 66% of
patients did not respond to thermal irrigation of the

70

Vertigo

Spontaneous nystagmus
~

~j3O"

0e.tokinetic nysta!l.mus
30%; left ~

~]30"

30%;right ~

60'Ysleft

6O'Ysright

~ ~

'OO%;left~
'OO%;right~
'20'Ysleft~

])Jo

~ ~3)JO
~ ~
~ ~j300

'20~right ~ ~ ~
Small field stimulation

60%;left~

60'Ysright~

~])Jo

~ ~
5 s

Fig.4.3. Original recordings of the horizontal component of spontaneous nystagmus and optokinetic nystagmus (OKN) 3,9, and 14
days after a right labyrinthine lesion. Directional preponderance of OKN toward the side of spontaneous nystagmus increases with
increasing stimulus speed and progressively vanishes with compensation for the vestibular im balance. (From Brandt et al. 1978.)

external auditory canal, and 34% showed reduced


responses. Two years later, however, 72% had normal
reactions, 12% showed reduced responses, and 16%
did not respond. They found complete recovery of
semicircular canal function in two-thirds of the
patients. In a more recent study, Okinaka et al. (1993)
found that caloric responses normalised in only 25
(42%) of 60 patients. Horizontal semicircular canal
paresis was found in about 90% 1 month after onset,
and in 80% after 6 months. The different results in
these two studies may be due to different criteria for
defining a pathological unilateral hyporesponsiveness. The prognosis for rare VN in children seems
better than in adults, since persistent canal paresis
was found in only 14% of 17 cases on reexamination
(Tahara et al. 1993).

MRimaging
Magnetic resonance imaging (MRI) has become
increasingly important for detecting labyrinthine or
eighth nerve disorders such as vestibular paroxysmia
(p. 117), Cogan's syndrome (p. 151), leptome,ngeal
carcinomatosis (p. 155), or meningitis. Due to recent
MRI advances it is now possible to demonstrate
facial nerve enhancement in Bell's palsy and
cochlear enhancement in sudden hearing loss.
However, to date attempts to image lesions of the
vestibular nerve or ganglion in patients with cryptogenie VN have failed (Hasuike et al. 1995; Strupp et
al. 1998c). None of 60 patients with idiopathie VN
exhibited contrast enhancement of the labyrinth,
vestibulocochlear nerve, or vestibular ganglion in

71

Vestibular neuritis

high-resolution MRI, even when high doses of


gadolinium (0.2 mmol!kg) were used (Strupp et al.
1998c).

Natural course
There is usually a sudden onset of the disease (sometimes preceded by a short vertigo attack hours or
days earlier) with rotatory vertigo, oscillopsia,
impaired fixation, postural imbalance, nausea, and
often vomiting. Patients feel severely ill and prefer to
stay immobilised in bed. They avoid head movements, which exaggerate symptoms, until the vertigo,
postural imbalance, and nausea subside, usually
after 1-3 days. After 3-5 days spontaneous
nystagmus is largely suppressed by fixation in the
primary position, although - depending on the
severity of the canal palsy - it is still present for 2-3
weeks with Frenzel's glasses and on lateral gaze
directed away from the lesion. After recovery of
peripheral vestibular function, in some patients
spontaneous nystagmus transiently reverses its
direction ("Erholungsnystagmus"), i.e. when the
centrally compensated lesion regains function.
"Erholungsnystagmus" then reftects a tone imbalance
secondary to compensation. Bechterew's phenomenon, areversal of post-unilaterallabyrinthectomy
spontaneous nystagmus occurring after contralateral
labyrinthectomy in animals (Bechterew 1883) or
humans (Zee et al. 1982; Katsarkas and Galiana
1984), is produced by a similar mechanism.
Stage assessment of VN is possible by means of
spontaneous and head-shaking nystagmus findings
(Matsuzaki and Kamei 1995). In the first stage spontaneous nystagmus of the paralytic type can be
observed after 4 weeks; subsequently head-shaking
nystagmus directed toward the unaffected ear indicates central compensation. Head-shaking nystagmus can disappear transiently during the process of
labyrinthine recovery or be directed toward the
affected ear as is spontaneous recovery nystagmus.
After 1-6 weeks most of the patients feel symptom free, even during slow body movements, but
actual recovery depends on whether and how quickly functional restitution of the vestibular nerve
occurs during "central compensation" and possibly
on how much physical exercise (p. 53) the patient
has done. Rapid head movements, however, may still
cause slight oscillopsia of the visual scene and
impaired balance for a second in those who do not
regain normallabyrinthine function (see following
paragraph). This explains why only 34 (57%) of 60
patients with VN reported complete relief from subjective symptoms at long-term follow-up (Okinaka
et al. 1993), roughly corresponding to the 50-70%

complete recovery rate of labyrinthine function


assessed by caloric irrigation (Meran and Pfaltz
1975; Okinaka et al. 1993; Ohbayashi et al. 1993).
Furthermore, in predisposed subjects the experience
of severe rotatory vertigo and imbalance in the acute
stage ofVN may initiate anxious introspection and
balance control which can escalate to panic attacks
and promote development of a phobic postural vertigo (p. 469).

High-frequency defect of VOR in permanent


peripheral vestibular lesion
It is possible to demonstrate a permanent, direction-

specific, high-frequency defect of the VOR in


patients whose semicircular canal function is not
restored. Ewald's second law (Ewald 1892; Hallpike
1961), which states that horizontal canal function
has a directional asymmetry, with ampullopetal
stimulation (cupula deftection toward the utricle)
being more effective than ampullofugal stimulation
(cupula deftection away from the utricle), predicts
this. Electrophysiological studies of primary vestibular afferents in the monkey during constant angular
accelerations have confirmed the law (Goldberg and
Fernandez 1971). Gain asymmetries have been
demonstrated in humans after acute unilateral
peripheral vestibular lesions, showing rotation
toward the side with the lesion (ampullopetal stimulation of the remaining intact labyrinth) resulted in
lower ga in than rotation away from the side with the
lesion (Baloh et al. 1977). In clinically compensated
unilaterallesions, Baloh et al. (1984) did not regularly
find significant asymmetry of gain during what they
called "high-frequency" sinusoidal rotation. But natural head movements with accelerations higher than
1000 0 /s 2 exceed by far the acceleration impulses provided by their rotary chair (140 0 /s 2 ). Unpredictable,
passive rotatory head impulses with accelerations up
to 4000 /s 2 demonstrated considerable asymmetries
in VOR gain even 1 year after a unilateral peripheral
vestibular lesion (Halmagyi et al. 1990).
Using a simple VOR-bedside test, Halmagyi and
Curthoys (1988) also demonstrated that there is no
central compensation of the directional asymmetry
of high-frequency canal function. When the head
was rapidly rotated toward the side with the lesion,
all 12 patients who had undergone unilateral
vestibular neurectomy made clinically evident,
oppositely directed, compensatory refixation saccades. This indicates a unilateral high-frequency
deficiency of VOR, produced by functional asymmetry of the remaining labyrinth. Furthermore,
the well-known clinical method of provoking
spontaneous nystagmus by passive head shaking
with Frenzel's glasses (Jung 1953; Kamei 1975)
0

72

reveals a unilateral labyrinthine loss even if it is


apparently compensated centrally. Hain et al. (1987)
were able to show that horizontal head shaking in
yaw elicits horizontal nystagmus with slow phases
that are initially directed toward the side of the
lesion and upward (fast phases directed toward the
unaffected ear). They assurne that "head-shaking
nystagmus is generated by the combination of a central velo city-storage mechanism, which perseverates
peripheral vestibular signals, and Ewald's second
law, which states that high-velocity vestibular excitatory inputs are more effective than inhibitory
inputs". Head-shaking nystagmus as a bedside test
allows not only clinical detection of a centrally compensated unilateral peripheral vestibular loss, but
also stage assessment of the spontaneous course of
VN (Matsuzaki and Kamei 1995).
Afferent cervical somatosensory input may substitute for absent vestibular input as part of central
vestibular compensation (p.60) after unilateral
peripheral vestibular deficit. A study of the perceptual
and oculomotor effects of neck muscle vibration in
VN (Strupp et al.1998a) showed the following:
l. there is an increase in muscle spindie input;
2. this increase is asymmetrical, restricted to the
affected side, and gradually builds up over
weeks;
3. the perceptual effects during vibration are secondary to changes in eye position rather than
changes in cortical representation of body orientation.

Thus, a unilateral increase in somatosensory


weight was demonstrated, which obviously substitutes for missing vestibular input.

Differential diagnosis
When based on careful his tory taking and clinical
evaluation, differential diagnosis is determined by
two elementary questions:
l. Is the clinical syndrome compatible with periph-

eral vestibular loss only or are there any central


neurological deficits incompatible with VN?
2. Are there any signs, symptoms, or clinical indications for a specific aetiology of an acute unilateral,
partial, or complete vestibular loss?
If central, there is only a small area in the lateral
medulla including the root entry zone of the vestibular nerve and the medial and superior vestibular
nuclei (see Central vestibular syndromes in yaw and
roll; p. 217), which is critical to avoid confusion with

Vertigo
peripheral vestibular nerve or labyrinthine lesions
(Brandt and Dieterich 1995). We have seen several
patients suffering from multiple sclerosis with pontomedullary plaques at the root entry zone of the
eighth nerve which mimicked VN (Fig.12.1; Brandt
et al. 1986; Dieterich and Bchele 1989). Small brainstern infarctions have also been reported to mimic
VN, e.g. in the form of a local brainstem syndrome
of rotatory vertigo with masseter paresis (Hopf
1987). Electrophysiological measures such as auditory
evoked potentials or masseter reflex and MRI (Hopf
1987; Francis et al. 1992) may help to identify brainstern disorders with few symptoms. The differential
diagnosis between central and peripheral causes of
unilateral vestibular loss is simple, if the patient presents with obvious additional brainstem signs. All
patients we observed with centrallesions mimicking
VN had incomplete horizontal semicircular canal
paresis and some additional ocular motor signs
(such as, saccadic vertical pursuit; direction-changing positional nystagmus) which were detected by
careful neuro-ophthalmological investigation. A central cause was always suspected prior to MRI. The
two disorders most relevant to the present discussion are multiple sclerosis and small brainstem
infarctions. Haemorrhages (cavernomas) or tumours
rarely manifest with purely acute rotatory vertigo
and horizontal semicircular canal paresis. Acoustic
neurinomas, which mostly arise from the vestibular
part of the eighth nerve, produce such a gradual
reduction in vestibular brainstem input from the
end-organ on the side of the tumour that central
compensation is capable of preventing vertigo.
However, acute rotatory vertigo and semicircular
canal paresis may rarely be the first manifestation of
a rapidly growing and larger tumour of the cerebellopontine angle. Then the critical site of the lesion is
peripheral, even though larger tumours compress
the brainstem and the flocculus.
The differential diagnosis of peripheral
labyrinthine and vestibular nerve disorders mimicking VN includes numerous rare conditions (see
Bilateral vestibulopathy, p. 127; Miscellaneous
vestibular nerve and labyrinthine disorders, p. 143).
Nevertheless, extensive laboratory examinations,
lumb ar puncture, and CT IMR imaging are not part
of the routine diagnostics of VN for two reasons: (1)
the rareness of these dis orders and (2) typical additional signs and symptoms indicative of other disorders. For example, the combination of vestibular
with auditory symptoms suggests herpes zoster oticus, if the ear is painful and blisters are observed in
the external auditory canal; or Cogan's syndrome
(p. 154), if inflammatory eye symptoms are found; or
Lyme borreliosis (Ishizaki et al. 1993), if the patient
reports arecent tick bite or an erythema migrans.

73

Vestibular neuritis

Thus, any further diagnostic procedures in patients


with VN are usually prompted and guided by the
unusual presentation of the syndrome, an atypical
course, or additional signs and symptoms. An initial
monosymptomatic vertigo attack in Meniere's disease or basilar migraine (p. 329) can be confused
with VN in a patient admitted to hospital in an acute
stage. The shortness of the attack and the patient's
rapid recovery, however, allow differentiation.

Aetiology and pathomechanism


In our dizziness unit VN is the third most common
cause of peripheral vestibular dis orders ({ 1) benign
paroxysmal positional vertigo, (2) Meniere's disease,
(3) VN) and accounts for about 5% of the patients.
Its usual age of onset is between 30 and 60 years
(Depondt 1973), and age distribution plateaus
between 40 and 50 years (Sekitani et al. 1993). There
is no significant sexual difference, although Meran
and Pfaltz (1975) found a peak for females in the
fourth decade and males in the sixth decade (Meran
and Pfaltz 1975). VN is relatively rare in children, but
it has repeatedly been reported to occur in children
aged 3-15 years, obviously affecting boys more frequently than girls (Tahara et al. 1993; Shirabe 1988).
VN in children seems to differ from VN in adults in
the following three aspects:
1. a higher frequency of preceding upper respir-

atory tract infections,


2. a more rapid recovery from vertigo and nystagmus,and
3. a better pro gnosis as to the recovery rate of
labyrinthine function assessed by follow-up
caloric testing (Shirabe 1988; Tahara et al. 1993;
Sekitani et al. 1993).

Pathomechanism
Normal vestibular end-organs generate an equal
resting-firing frequency which is the same on both
sides. This continuous excitation (resting dis charge
rate in monkey :::: 100 Hz, Goldberg and Fernandez
1971; 18000 vestibular afferents for each labyrinth,
Bergstrom 1973) is transmitted to the vestibular
nuclei via vestibular nerves. Pathological processes
affecting an end-organ alter its firing frequency,
thereby creating a tone imbalance. This imbalance
causes most of the manifestations of the vertigo
syndrome: perceptual, ocular motor, postural, and
vegetative (nausea).

As distinct from bilateral vestibulopathy (p. 127),


unilateral semicircular canal paresis can be largely
substituted for by the redundant canal input from
the unaffected labyrinth. Angular head accelerations
are detected by three pairs of semicircular canals
and linear head accelerations by two pairs of
otoliths. These sensors induce compensatory eye
movements (slow phases) in the opposite direction
to head acceleration and transduce the sensation of
head motion. Sensorimotor transformation proceeds via canal planes to planes of eye movements so
that the neurons always contact their two respective
extraocular muscles. This means that alesion of a
single semicircular canal induces a spontaneous nystagmus with slow phases in the "off-direction" of
that canal. If multiple canals are lesioned, the slow
phases should be in a direction that is a weighted
vector sum of the axes of the involved canals (Fetter
and Dichgans 1996). The direction of head rotation
is sensed by corresponding on-and-off modulation
of the resting activity (on: 100-500 Hz; off: 100-0
Hz) of the right and left canals, corroborating in
pairs for the particular plane of motion (yaw = horizontal semicircular canals, right and left). Loss of
function in the on-direction (head rotation to the
right with right horizontal semicircular canal paresis) is still sensed by the opposite canal, wh ich is
stimulated (inhibited) in its off direction.
Modulation of the neuronal activity in the off direction is limited, and as the speed of head rotations
increases, the firing rate of the neurons will reach
zero; this is also called Ewald's second law (Ewald
1892) (see also High-frequency defect ofVOR, p. 71).
Vertigo, spontaneous nystagmus with oscillopsia,
and postural imbalance in VN are the appropriate
stimuli to promote central vestibular compensation
(p. 55) and vestibular substitution by visual and
somatosensory input (p. 60). Since vestibular compensation is less perfeet than generally believed, especially for dynamic conditions, further mechanisms
such as sensory substitution by vision or proprioception in part replace the missing vestibular input.
There is, for example, a measurably increased influence of cervical proprioception on spatial orientation and gaze in space ipsilateral to a peripheral
vestibular lesion (Strupp et al. 1998b). Vestibular
exercises (p. 76) and pharmacological substances
(p. 58) may facilitate these processes.

Vestibular neuritis - a partial unilateral


vestibular loss
Does VN produce a complete or a partial unilateral
vestibular paralysis? The coexistence of a complete
VN and benign paroxysmal positioning vertigo

Vertigo

74

(p. 251) in the same individual, in the same ear, at


the same time seems impossible, for this implies the
simultaneous function and loss of function of one
labyrinthine structure (Fig. 4.4). The repeated clinical observation of this apparently paradoxical coincidence led us to draw the following conclusions
(Bchele and Brandt 1988).
VN typically affects only part of the vestibular
nerve trunk, usually the superior division (horizontal and anterior semicircular canals, maculae of the
utricle and anterosuperior part of the saccule),
which has its own path and ganglion (Lorente de No
1933; Sando et al. 1972), whereas the inferior part
(posterior semicircular canal) is spared (Fig. 4.5).
This hypothesis of partial involvement of the
vestibular nerve is supported by findings of temporal bone pathology (Schuknecht and Kitamura
1981) and also by the histopathology of a case of
herpes zoster oticus (Proctor et al. 1979). In the latter
case the otolith apparatus and the posterior semicircular canal remained intact. The earlier hypothesis
of Lindsay and Hemenway (1956), on the other hand,
convincingly explained the coexistence of VN and
benign paroxysmal positioning vertigo as due to a
vascular pathogenesis (Fig. 4.4). If an ischaemic
event involves only the anterior vestibular artery, it

would cause a contraversive horizontal nystagmus in


the acute stage but no ipsilateral response to thermic
irrigation of the horizontal canal; it could also promote cupulolithiasis of otoconia by ischaemic
degeneration of the utricular macula. Histological
investigations by these authors found degeneration
of the nerve fibres leading to the horizontal and the
anterior semicircular canals and the utricle, whereas
the posterior ampullary nerve appeared intact.
Schuknecht and Kitamura (1981) favoured a viral
aetiology and, in order to disprove a vascular aetiology, even tried to declare that Lindsay and
Hemenway's case was of viral origin. Recently, analysis
of 3D properties of the vestibular ocular reflex in
patients with VN clearly demonstrated that the vectors of the spontaneous nystagmus clustered
between the expected directions for lesions of either
the horizontal or a combined lesion of the horizontal
plus the anterior semicircular canals (Fetter and
Dichgans 1996). These data strongly support lesion
of the superior division of the vestibular nerve, sparing the inferior division. Bhmer et al. (1997), however, doubt that 3D analysis of spontaneous
nystagmus allows accurate localisation of a peripheral vestibular lesion. Involvement of the inferior
division of the vestibular nerve is possible in some

Spontaneous nystagmus (eyes closedl

20
hOrlz

ENG 20.

-t-.,.....,.-~"""''''''-''~-v-J'V"'''''~ "-"N'-I ~~
L

~~
lett aar 44 C

right ear 44 C

lelt ear 30 C

right ear 30 C

~~

Poslt/onal nystagmus lelt

vert

ENG

3rd

Fig.4.4. Patient with vestibular neuritis and benign paroxysmal positioning vertigo simultaneously in the same (left) ear, suggesting
partial rather than complete vestibular para lysis. Original recording of spontaneous nystagmus (top), caloric testing (middle) and the
vertical component of paroxysmal nystagmus precipitated by rapid positioning manoeuvre toward the left affected ear (bottom).
There is still a spontaneous nystagmus beating to the right; thermic irrigation reveals unresponsiveness of the left horizontal semicircular canal: positioning nystagmus typically fatigues during the repetitive three positioning manoeuvres. (From Bchele and Brandt
1988.)

75

Vestibular neuritis

Fig.4.5.

Vestibular neuritis, a partiallabyrinthine lesion with


horizontal semicircular canal paresis: vascular or viral aetiology? A
theoretical explanation is that only the anterior vestibular artery
(AVA) is affected, sparing the posterior branch which supplies the
posterior canal (top). The more likely explanation is a viral aetiology
affecting parts of the vestibular nerve (VN), in particular the horizontal ampullary nerve, but sparing the inferior division, the functioning of which is necessary for the simultaneous occurrence of
vestibular neuritis and benign paroxysmal positioning nystagmus
in the same ear. AC: anterior canal; HC: horizontal canal; pe: posterior canal.

cases of VN, and the absence of vestibular evoked


myogenic potentials (saccular origin) may indicate
this (Murofushi et al. 1996). The occurrence of a
complete ocular tiIt reaction indicates involvement
of both superior and inferior parts of the vestibular
nerve (Arbusow et al. 1998).

Viral or vascular aetiology?


Historical discussion

In the past, two main causes were proposed: either


inftammation of the vestibular nerve (Ruttin 1909;
Nylen 1924; Hallpike 1949; Dix and Hallpike 1952;
Aschan and Stahle 1956) or vascular disturbance,
which could be due to labyrinthine ischaemia
(Lindsay and Hemenway 1956) or even microvascular disturbances caused by infection (Meran and
PfaItz 1975). There was some doubt about the existence of aseparate diagnostic entity, and acute
vestibular paralysis was considered to be merely a
symptom of either labyrinthine ischaemia or

inftammation of the vestibular nerve (Kornhuber


and Waldecker 1958).
The histological findings in single cases reported
by Hilding et al. (1968) suggest infectious pathogenesis. On the basis of a few autopsy studies, in which
the pathological findings were similar to those
occurring with known viral disorders, Schuknecht
and Kitamura (1981) deduced that typical VN is in
fact a viral neuritis of the superior vestibular nerve
trunk. Schuknecht (1985) distinguishes between
acute viral labyrinthitis (cochlear and/or vestibular),
acute viral neuritis, and delayed endolymphatic
hydrops as the sequel to labyrinthitis. Similarly, viral
cochleitis is a convincing explanation for sudden
idiopathic sensorineural hearing loss (without vertigo), a conclusion supported by findings of temporal
bone pathology (Schuknecht and Donovan 1986).
Finally, acute bilateral sequential VN has been
described by Schuknecht and Witt (1985) and by
Ogata et al. (1993). This condition has a poor prognosis, with permanent but somewhat lessening disequilibrium arising from a bilateral partialloss of
vestibular function (p. 127). Herpes zoster oticus
(p. 146) is considered an entity separate from VN if
it manifests with auditory and vestibular symptoms
(Hunt 1908; Proctor et al. 1979; Longridge 1989). The
mumps virus can cause not only deafness but also
vertigo and impairment of caloric responses (Hyden
et al. 1979).
Arguments for viral aetiology

The most popular theory is that of viral aetiology,


but the evidence for it remains circumstantial
(Nadol1995; Tran Ba Huy 1994). The following arguments are presented to support a viral aetiology:
1. VN shows an epidemic occurrence in certain

periods of the year, and there is a high frequency


of preceding or concurrent upper respiratory
tract infections (about 30% in adults, Silvoniemi
1988; Sekitani et al. 1993); however, a critical
appraisal of the significance of these epidemiological arguments seems called for (Tran Ba Huy
1994).
2. Vestibular nerve histopathology in cases of VN
(Schuknecht and Kitamura 1981) is similar to
that seen in single cases of herpes zoster oticus,
when temporal bone histopathology was available (Zajtchuk et al. 1972). Temporal bone specimens from three patients with a history of
chronic recurrent vertigo of unknown cause
showed varying degrees of inftammation and
destruction in the vestibular system, and mild
involvement of the cochlear system consistent

Vertigo

76

with postinfectious inflammatory changes


(Ishiyama et al. 1997).
3. Cerebrospinal fluid investigations show an
increase in protein (not in cells) beginning about
2 weeks after onset of VN. This could be due to
increased entry of plasma pro teins caused by
either a disruption of the blood-brain barrier or
local immunoglobulin production (rising antibody titres) or to demyelination of the vestibular
nerve (Matsuo 1986; Matsuo et al. 1989).
Increases in various serum antibody titres have
been found in about one-half of patients with
VN (Shimizu et al. 1993; Hirata et al. 1989), but
no increase in IgG or viral antibody titres in the
cerebrospinal fluid (Matsuo et al. 1989).
4. Herpes simplex virus (HSV) DNA was repeatedly
detected in autopsied human vestibular ganglia
by using polymerase chain re action (PCR)
(Furuta et al. 1993). This indieates that vestibular
ganglia are latently infected by HSV-l; however,
the latency-associated transcript was found to be
negative.
5. An animal model ofVN was developed by inoculating herpes simplex virus type 1 into the auride
of miee (Hirata et al. 1995). Postural deviation
was observed in 5% of the mice 6-8 days after
the inoculation. Degeneration of Scarpa's ganglion and HSV-l antigens were found only in
symptomatic animals. Vestibular symptomatology can be induced by intraperitoneal, intracerebral, intralabyrinthine, or intracutaneous
inoculation of various viral agents (Davis and
Johnson 1976; Davis 1993; Hirata et al. 1995).
The sum of all these arguments fails to establish a
common aetiology and pathomechanism for VN,
and does not identify a single or typieal causative
virus. If herpes simplex is the most likely candidate,
it can be assumed to reside in a latent state in the
vestibular ganglia, e.g. in the ganglionie nudei as has
been reported in other cranial nerves (Nahmias and
Roizman 1973). All combinations ofHSV-l positive
and negative genieulate, vestibular, and spiral ganglia were found in 18 human temporal bones (Schulz
et al. 1998). As a result of intercurrent factors, it suddenly replicates and induces an autoimmune reaction, leading to inflammation and oedema, and
subsequent demyelination of the nerve (Tran Ba Huy
1994), whieh increases protein in cerebrospinal fluid,
indieating a deficient barrier between blood and
cerebrospinal fluid (Matsuo et al. 1989). Increases in
serum antibody titres, however, were found not only
for herpes simplex, but also for Epstein-Barr, rubella,
adeno, influenza, and cytomegaloviruses. Both
convergent and divergent data in support of viral

aetiology must be critieally weighed against each


other, partieularly with respect to the therapeutie
consequences.

Site of the lesion


It is commonly accepted that typieal VN is due to a
lesion of the superior division of the vestibular
nerve (Schuknecht and Kitamura 1981; Bchele and
Brandt 1988; Fetter and Diehgans 1996). But textbooks and artieies may still cause confusion, for
some reports attribute the disorder also to
labyrinthine (Bergenius and Borg 1983), brainstem
(Wennmo and Pyykk 1982; Hopf 1987), or cerebellar structures (Kmpf 1986). These authors support
their arguments with findings of associated elevated
stapedius reflex thresholds (Bergenius and Borg
1983; Bagger-Sjoback et al. 1993), central oculographie patterns (Wennmo and Pyykk 1982; Imate
et al. 1995), paresis of the masseter musde with
impairment of the masseter reflex (Hopf 1987), or
cerebellar infarction in the CT or MRI scans (Kmpf
1986; Magnusson and Norrving 1993). All these conditions, in partieular MS plaques of the root entry
zone of the eighth nerve, differ from typical VN.
Most importantly, dinieal examination usually disdoses their difference by detecting additional symptoms beyond those common for typieal VN (see
Differential diagnosis, p. 72).
Thus, all forms of acute unilateral peripheral
vestibular dysfunction can mimie the dinieal
appearance and spontaneous course ofVN, especially
if other symptoms are lacking. The dysfunction can
then occur anywhere from the labyrinth to the
vestibular nudei.

Management
Management ofVN involves
1. medieal treatment

antivertiginous drugs (dimenhydrinate, scopolamine) or benzodiazepines


cortieosteroids
2. physieal therapy (vestibular exercises)
During the first 1-3 days, when nausea is pronounced, benzodiazepines or vestibular sedatives
such as antihistamine dimenhydrinate (Dramamine )
50-100 mg every 6 h or the "antieholinergie" scopolamine (Transderm-Scop) 0.6 mg, can be administered parenterally for symptomatic relief; the major

Vestibular neuritis

side effect is general sedation (Brandt et al. 1974).


Transdermal application of scopolamine hydrobromide avoids some of the side effects of conventional
me ans of administration. The most probable sites of
primary action are the synapses of the vestibular
nuclei, which exhibit a reduced discharge and dirn inished neuronal re action to body rotation. These
drugs should not be given after nausea disappears,
because they prolong the time required to achieve
central compensation (p. 58).
Animal models have shown that glucocorticoids
facilitate central vestibular compensation (Jerram et
al. 1995; Yamanaka et al. 1995). To date two studies
on patients have reported a beneficial effect of corticosteroids on the course ofVN. The study by Ariyasu
et al. (1990) included 20 randomly selected patients
and was double-blind, prospective, placebocontrolled, and crossover; however, uncertainties
about the precise diagnosis remained ("acute
vestibular vertigo"). Patients took a single 32 mg
dose of methylprednisolone orallyon the first dayand
divided doses of 16 mg twice a day for 3 days; then
the dosage was tapered to zero after 8 days. While
the second study by Ohbayashi et al. (1993) had no
clearly prospective design, it also reported that corticosteroids facilitated early recovery from vertigo and
nystagmus. The recovery rate of caloric responses at
follow-up was significant for moderate horizontal
semicircular canal paresis but not for marked
paresis. The administration of steroids included
infusion of hydrocortisone (initial dosage of 500 mg
was decreased gradually by 100 mgl2 days for 10
days) or oral prednisolone starting with 30-40
mg/day. Antiviral substances, such as acyclovir, have
not yet been systematically studied. A double-blind
study conducted by Adour et al. (1990) found that
acyclovir-prednisone is superior to prednisone alone
for treating Bell's palsy.
Another mode of treatment (which should complement medical therapy) is physical therapy with
the Cawthorne-Cooksey (Cawthorne 1944) exercises,
modified according to current knowledge of vestibular physiology (Table 4.1; Brandt 1986), also called
vestibular rehabilitation (Herdman 1994; Foster
1994). Vestibular exercises consist mainly of eye,
head, and body movements designed to provoke a
sensory mismatch; they enhance compensation by
facilitating central recalibration, although the symptoms are initially uncomfortable. Therapy for
vestibular imbalance should be designed to expose
the patient increasingly to unstable body positions
in order to facilitate rearrangement and recruitment
of control capacities (Brandt et al. 1981) (see
Vestibular exercises, p. 52). Elderly patients seem to
take longer to recover (Sloane et al. 1989; Ishikawa et
al. 1993). The experimentally observed differential

77
Table 4.1.

Physical therapy for acute, unilaterallabyrinthine lesions

Clinical stage

Physical exercise

I. Approximately days 1-3

Nausea
Spontaneous
nystagmus with
fixation

No exercise; bed rest


Head immobilisation

Eyes closed

Strategy
Prevent falls
Avoid active head
accelerations leading to
"cross-coupled" effects
Avoid visual-vestibular
mismatch

11. Approximately days 3-5

No spontaneous
nausea

Exercise in bed (supine


and sitting) with rapid
mobilisation
Incomplete suppression 1. Fixation straight
of spontaneous
ahead; voluntary
nystagmus by
saccades and eccentric
fixation straight ahead gaze-holding (10,
20, and 40 horizontall
vertical); reading
exercise
Smooth pursuit
(finger movements or
pendulum 20-40;
20-60 /s)

Visual control of
stabilisation of gaze
in space by suppressing
spontaneous
nystagmus through
voluntary fixation
impulse (retinal slip)
Visually guided
controlof
target fixation

Active head oscillations


with fixation of a
stationary target
at I m distance
(0.5-2 Hz; 20-30;
yaw> pitch > roll)
2. First balance
exercise-free sitting,
stance, and guided
gait (eyes open, eyes
closed)

Provoke vestibular
stimuli for
recalibration of VOR
under visual control
of retinal slip of the
viewed target
Circulatory training,
prophylaxis of
thrombosis

111. Approximately days 5-7

Suppression of
spontaneous
nystagmus with
fixation straight
ahead, but
continued gaze
nystagmus in the
direction of fast
phase,and
spontaneous
nystagmus with
Frenzel's glasses

1. Static stabilisation: Recalibrate


Four point stance
visuovestibuloStance on one knee
spinal reflexes for
and onefoot
postural control and
Upright stance (eyes
eye-head
open/eyes closed;
coordination du ring
head upright/head
free body
extension)
movements
2. Dynamic stabilisation:
Smooth pursuit and
head oscillation exercises
during free stance as
described in preceding
section
Exercises with rope, ball,
and club under fixation
(eye and head) ofthe
instrument (sittingl
standing/walking)

IV. Approximately weeks 2-3

No spontaneous
vertigo
Weak spontaneous
nystagmus with
Frenzel's glasses

Complex balance
exercise, successive
increase in difficulty,
above the demands for
postural control under
daily living
conditions

Expose the
subjectively"recovered
patient" increasingly
to unstable body
positions in order to
facilitate
rearrangement and
recruitment of control
capacities

Vertigo

78

effect of monocular versus binocular vision on


vestibular compensation remains uncertain and
unclear (Ishikawa and Togawa 1988; Hamid et al.
1991).
Animal experiments have shown that visual and
physical exercises promote central compensation of
spontaneous nystagmus (Courjou et al. 1977) as wen
as postural reflexes in locomotion (Igarashi et al.
1981; Lacour 1984).A prospective study (Strupp et
al. 1998b) was able to demonstrate that specific
vestibular exercises significantly improve vestibulospinal compensation in patients with acute VN
(Fig. 4.6). Central compensation of unilateral peripheral vestibular lesions involves multiple processes
occurring in distributed networks at different
locations (spinal cord, vestibular nuclei, commissural brainstem connections between vestibular nuclei)

and with different time courses (see Vestibular compensation, p. 55). Pharmacological and metabolie
studies in animals suggest that the state of central
compensation for peripheral vestibular lesions is
both dynamic and fragile (Zee 1985). Alcohol, phenobarbital, chlorpromazine, diazepam, Ca2+ -channel
antagonists (Darlington and Smith 1992), and
ACTH-antagonists (Gilchrist et al. 1990) may retard
compensation; caffeine, amphetamin es, and
ACTH accelerate compensation; cholinomimetics,
cholinesterase inhibitors, adrenergic agents, GABAagonists, and alcohol can (re)produce decompensation. It still remains to be proven if the use of drugs
accelerates compensation in patients (Smith and
Darlington 1994) (p. 58).
Table 4.2 summarises the information given in
this chapter about VN.

45
40
35
;:::

30

--S

25

-5

20

control

D vestibular exercises
p

< 0.001

'-'
~

0-

>.
c.;
~

CZl

, - - - - - - - - - -_ _ n = 20

15
10

...................................

......................... .

.fIf----

yn =

19

0
0

10

15

20

25

30

time (day after symptom onset)


Fig.4.6. Efficacy of specific vestibular exercises for postural sway in patients with acute vestibular neuritis_ Time course of the
changes in total sway path values (SP) for the controls (n = 20) and physiotherapy (n = 19) groups: vestibular exercises improved central vestibulospinal compensation_ For postural control SP values (m/min, mean SD) were measured in patients with eyes closed and
standing on a compliant foam-padded posturography platform_ The total SP is the length of the path described by the centre of force
during a given time (20 s), which is generated by the inherent instability of a subject standing on a recording platform. SP is approximated by the sum of the distances between two consecutive sampling points in the antero-posterior (sagittal = x) plane, Le_ sagittal
sway (calculated as II ~xll, mediolateral (frontal = y) plane, Le. frontal sway (calculated as II~yll, or for two dimensions as - the total SP
- (calculated as I'!q ~x21+I l~y21)). There was a significant difference (ANOVA, p < 0.001) between the two groups at the statistical
end point (day 30 after symptom onset). The dotted line indicates the normal range. (From Strupp et al. 1998b.)

Vestibular neuritis
Table 4.2. Vestibular neuritis
Clinical syndrome
- Acute onset of sustained
- rotatory vertigo
- postural imbalance with falls toward the affected ear
- horizontal-rotatory spontaneous nystagmus (toward the
unaffected earl
- nausea and vomiting
- unilateral hypo- or unresponsiveness in caloric testing
Incidence/age/sex
Third most common cause of peripheral vestibular vertigo that
manifests throughout life (affects mainly ages 30 to 60 years; rare in
children) without preference of sex
Pathomechanism
Acute partial unilateralloss of labyrinthine function (horizontal and
anterior semicircular canal paresis) with a vestibular tone imbalance in
yaw and roll planes
Aetiology
Most probably viral infection of the superior division of the vestibular
nerve trunk
Course/prognosis
Spontaneous recovery within 1-6 weeks due to
- (contralateral) vestibular, somatosensory, and visual substitution
of the vestibular deficit
- central compensation of vestibular tone imbalance
- peripheral restoration of labyrinthine function (incomplete in
about 50%)
Better prognosis and higher recovery rate in children
Management
Medical treatment
- antivertiginous drugs (dimenhydrinate, scopolamine)
- corticosteroids
Physical therapy (vestibular exercises)
Differential diagnosis
- Acute central brainstem lesions at the root entry zone of the eighth
nerve and the vestibular nucleus (MS plaques, small pontomedullary
infarcts)
- Peripherallabyrinthine and vestibular nerve disorders, e.g. vascular
(AICA infarcts), inflammatory (Lyme borreliosisl, or immunological
(Cogan's syndrome) disorders

References
Adour KK, Ruboyianes JM, Von Doersten PG, Byl FM, Trent CS,
Quesenberry CP Jr, Hitchcock T (1990) The beneficial effect of
methylprednisolone in acute vestibular vertigo. Arch
Otolaryngol Head Neck Surg 116:700-703
Arbusow V, Dieterich M, Strupp M, Dreher A, Jger L, Brandt T
(1998) Herpes zoster neuritis involving superior and anterior
parts of the vestibular nerve causes ocular tilt re action. Neuroophthalmology 19:17-22
Ariyasu L, Byl FM, Sprague MS, Ardour KK (1990) The beneficial
effect of methylprednisolone in acute vestibular vertigo. Arch
Otolaryngol Head Neck Surg 116:700-703
Aschan G, Stahle J (1956) Vestibular neuritis. J Laryngol
70:497-511
Bagger-Sjbck D, Perols 0, Bergenius J (1993) Audiovestibular
findings in patients with vestibular neuritis: a long-term followup study. Acta Otolaryngol (Stockh) Suppl 503: 16-17

79
Baloh RW, Honrubia V, Konrad HR (1977) Ewald's second law reevaluated. Acta Otolaryngol (Stockh) 83:475-479
Baloh RW, Honrubia V, Yee RD, Hess K (1984) Changes in the
human vestibulo-ocular reflex after loss of peripheral sensitivity.Ann NeuroI16:222-228
BechterewW (1883) Ergebnisse der Durchschneidung des Nervus
acusticus nebst Errterung der Bedeutung der semicirculren
Kanle fr das Gleichgewicht. Pflgers Arch ges Physiol
30:312-347
Bergenius J, Borg E (1983) Audio-vestibular findings in patients
with vestibular neuritis. Acta Otolaryngol (Stockh) 96:389-395
Bergstrom B (1973) Morphology of the vestibular nerve H. The
number of myelinated vestibular nerve fibres in man at various
ages.Acta Otolaryngol (Stockh) 76:173-179
Bhmer A, Rickenmann J (1995) The subjective visual vertical as a
clinical parameter of vestibular function in peripheral vestibular diseases. J Vestib Res 5:33-45
Bhmer A, Straumann D, Fetter M (1997) Three-dimensional
analysis of spontaneous nystagmus in peripheral vestibular
lesions.Ann Otol Rhinol LaryngoI106:61-68
Brandt Th, Daroff RB (1980) The multisensory physiological and
pathological vertigo syndromes. Ann NeuroI7:195-203
Brandt Th, Dieterich M (1995) Central vestibular syndromes in
roll, pitch and yaw planes: Topographic diagnosis of brainstem
disorders. Neuro-ophthalmology 15:291-303
Brandt Th, Dichgans J, Wagner W (1974) Drug effectiveness on
experimental optokinetic and vestibular motion sickness.
Aerospace Med 45:1291-1297
Brandt Th, Allum JHJ, Dichgans J (1978) Computer analysis of
optokinetic nystagmus in patients with spontaneous nystagmus
of peripheral vestibular origin. Acta Otolaryngol (Stockh) 86:
115- 122
Brandt Th, Krafczyk S, Malsbenden I (1981) Postural imbalance
with head extension: improvement by training as a model for
ataxia therapy. Ann NY Acad Sci 374:646-649
Brandt Th, Dieterich M, Bchele W (1986) Postural abnormalities
in central vestibular brainstem lesions. In: Bles W, Brandt Th
(eds) Disorders of posture and gait. Elsevier, Amsterdam, pp
142-156
Bchele W, Brandt Th (1988) Vestibular neuritis, a horizontal
semicircular canal paresis? Adv Oto-Rhino- Laryngol
42:157-161
Cawthorne T (1944) The physiologic basis for head exercises. J
Chart Soc Physiother 106-107
Courjou JH, Jeannerod M, Ossuzio I, Schmidt R (1977) The role of
vision on compensation of vestibulo-ocular reflex after hemilabyrinthectomy in the cat. Exp Brain Res 28:235-248
Darlington Cl, Smith PF (1992) Pre-treatment with a Ca2+ channel antagonist facilitates vestibular compensation. NeuroReport
3:143-145
Davis LE (1993) Viruses and vestibular neuritis: review ofhuman
and animal studies. Acta Otolaryngol (Stockh) Suppl
503:700-773
Davis LE, Johnson RT (1976) Experimental viral infections of the
inner ear. I. acute infections of the newborn hamster labyrinth.
Lab luvest 34:349-356
Depondt M (1973) La neuronite vestibulaire, paralysie vestibulaire
11 caracteres particulier. Acta Oto- Rhino- Laryngol Belg
27:323-359
Dieterich M, Bchele W (1989) MRI findings in lesions at the
entry zone of the eighth nerve. Acta Otolaryngol (Stockh) Suppl
468:385-389
Dix MR, Hallpike CS (1952) The pathology symptomatology and
diagnosis of certain common dis orders of the vestibular
system.Ann Otol (St Louis) 61:987
Ewald R (1892) Physiologische Untersuchungen ber das
Endorgan des Nervus octavus. Bergmann, Wiesbaden
Fetter M, Dichgans J (1996) Vestibular neuritis spares the inferior
division of the vestibular nerve. Brain 119:755-763

80
Foster CA (1994) Vestibular rehabilitation. Baillieres Clin Neurol
3:577-592
Francis DA, Bronstein AM, Rudge P, du Boulay EPGH (1992) The
site of brainstem lesions causing semieircular canal paresis: an
MRI study. J Neurol Neurosurg Psychiatry 55:446-449
Furuta Y, Takasu T, Fukuda S, Inuyama Y, Sato KC, Nagashima K
(1993) Latent herpes simplex virus type I in human vestibular
ganglia. Acta Otolaryngol (Stockh) Suppl 503:85-89
Gilchrist DP, Smith PF, Darlington CL (1990) ACTH (4-10) accelerates ocular motor recovery in the guinea pig following
vestibular deafferentation. Neurosei Lett 118:14-16
Goldberg JM, Fernandez C (1971) Physiology of peripheral neurons innervating semieircular canals of the squirrel monkey: I.
Resting discharge and response to constant angular accelerations. J Neurophysiol 34:635-660
Hain TC, Fetter M, Zee DS (1987) Head-shaking nystagmus in
patients with unilateral peripheral vestibular lesions. Am J
OtolaryngoI8:36-47
Hallpike CS (1949) The pathology and differential diagnosis of
aural vertigo. Proc 4th Intern Congress Otolaryngol, London, Br
Med Ass 2:514
Hallpike CS (1961) On the case for repeal of Ewald's second law:
Some introductory remarks. Acta Otolaryngol (Stockh) Suppl
149:7-14
Halmagyi GM, Curthoys IS (1988) A clinical sign of canal paresis.
Arch NeuroI45:737-739
Halmagyi GM, Curthoys IS, Cremer PD, Henderson CJ, Todd MJ,
Staples MJ, D'Cruz DM (1990) The human horizontal vestibuloocular reflex in response to high-acceleration stimulation
before and after unilateral vestibular neurectomy. Exp Brain
Res 81:479- 490
Hamid M, Roberts VI, Haddon K (1991) Monocular and binocular
suppression of vestibular nystagmus. Acta Otolaryngol
(Stockh) SuppI481:424-427
Hasuike K, Sekitani T, Imate Y (1995) Enhanced MRI in patients
with vestibular neuronitis. Acta Otolaryngol (Stockh) Suppl
519:272-274
Herdman SJ (1994) Vestibular rehabilitation. FA Davis,
Philadelphia
Hilding DA, Kanda T, House WF (1968) Vestibular neuronitis and
small acoustic neuroma: electron microscopic observations.
Otol Clin N Am 305-318
Hirata Y, Sekitani T, Okinaka Y, Matsuda Y (1989) Serovirological
study of vestibular neuronitis. Acta Otolaryngol (Stockh) Suppl
468:371-373
Hirata Y, Gyo K, Yanagihara N (1995) Herpetic vestibular neuritis:
an experimental study. Acta Otolaryngol (Stockh) Suppl
519:93-96
Hopf HC (1987) Vertigo and masseter paresis. A new local brainstern syndrome probably of vascular origin. J NeuroI235:42-45
Hunt JR (1908) A further contribution to the herpetic inflammation of the geniculate ganglion. Am J Med Sei 136:226 - 241
Hyden D, dkvist LM, Kylen P (1979) Vestibular symptoms in
mumps deafness.Acta Otolaryngol (Stockh) SuppI360:182-183
Igarashi M, Levy JK, 0-Uchi T, Reschke MF (1981) Further study
of physical exereise and locomotor balance compensation after
unilateral labyrinthectomy in squirrel monkey. Acta
Otolaryngol (Stockh) 92:101-105
Imate Y, Sekitani T, Okami M, Miura M (1995) Central disorders in
vestibular neuronitis. Acta Otolaryngol (Stockh) Suppl
519:204-205
Ishikawa K, Togawa K (1988) Effect of blindfolding one eye on
vestibular compensation in guinea pigs. Acta Otolaryngol
(Stockh) Supp1198; 47:55-60
Ishikawa K, Edo M, Togawa K (1993) Clinical observations of 32
cases of vestibular neuronitis. Acta Otolaryngol (Stockh) Suppl
503:13-15
Ishiyama A, Ishiyama GP, Lopez I, Eversole LR, Honrubia V, Baloh
RW (1997) Histopathology of idiopathic chronic recurrent vertigo. Laryngoscope 106: 1340 -1346

Vertigo
Ishizaki H, Pyykk I, Nozue M (1993) Neuroborreliosis in the aetiology of vestibular neuronitis. Acta Otolaryngol (Stockh) Suppl
503:67-69
Jerram AH, Darlington CL, Smith PF (1995) Methylprednisolone
reduces spontaneous nystagmus following unilateral
labyrinthectomy in guinea pig. Eur J PharmacoI275:291-293
Jung R (1953) Nystagmographie. Zur Physiologie und Pathologie
des optisch-vestibulren Systems beim Menschen. In: von
Bergmann G, Frey W, Schwieck H (eds) Handbuch der Inneren
Medizin, 4th Ed. Springer, Berlin, Vol511, pp 1325-1379
Jung R, Mittermaier K (1939) Zur objektiven Registrierung und
Analyse verschiedener Nystagmusformen: Vestibulrer,
optokinetischer und spontaner Nystagmus in ihren
Wechselbeziehungen. Arch Ohr Nas Kehlkopfheilk 146:410-439
Kamei T (1975) Der biphasisch auftretende Kopfschttelnystagmus Arch Otolaryngol 209: 59-67
Katsarkas A, Galiana HL (1984) Bechterew's phenomenon in
humans. A new explanation. Acta Otolaryngol (Stockh) Suppl
406:95-100
Kmpf, D (1986) Der benigne pseudovestibulre Kleinhirninsult.
Nervenarzt 57:163-166
Kornhuber H, Waldecker G (1958) Akute isolierte periphere
Vestibularisstrungen. Arch Ohr usw Heilk Z Hals usw Heilk
173:340-346
Lacour M (1984) Reapprentissage et periode postoperatoire sensible dans la restauration des fonctions nerveuses. Exemple de la
compensation vestibulaire et implications cliniques. Ann OtoLaryng 101:177-187
Lindsay JR, Hemenway WG (1956) Postural vertigo due to unilateral sudden partialloss of vestibular function. Ann Otolaryngol
65:692-706
Longridge NS (1989) Recurrent vestibulopathy: Support for a viral
aetiology. J Otolaryngol18: 99- 100
Lorente De No R (1933) Vestibulo-ocular reflex arc. Arch Neurol
Psychiat 30:245-291
Magnusson M, Norrving B (1993) Cerebellar infarctions and
"vestibular neuronitis". Acta Otolaryngol (Stockh) Suppl
503:64-66
Matsou T (1986) Vestibular neuronitis: serum and CSF virus antibody titre. Auris Nasus Larynx 13: 111-134
Matsuo T, Sekitani T, Honjo S, Imate Y, Inokuma T (1989)
Vestibular neuronitis. Pathogenesis in the view of virological
study of CSF. Acta Otolaryngol (Stockh) SuppI468:365-369
Matsuzaki M, Kamei T (1995) Stage assessment of the progress of
continuous vertigo of peripheral origin by means of spontaneous and head-shaking nystagmus findings. Acta
Otolaryngol (Stockh) SuppI519:188-190
Meran A, Pfaltz CR (1975) Der akute Vestibularisausfall. Arch OtoRhino-LaryngoI209:229-244
Murofushi T, Halmagyi GM, Yavor RA, Colebatch JG (1996) Absent
vestibular evoked myogenic potentials in vestibular neurolabyrinthitis: An indicator of inferior vestibular nerve involvement? Arch Otolaryngol Head Neck Surg 122:845-848
Nadol JB (1995) Vestibular neuritis. Otolaryngol Head Neck Surg
112:162-172
Nahmias AI, Roizman BC (1973) Infection with herpes-simplex
viruses land 2. N Engl J Med 289:719-725
Nylen CO (1924) Some cases of ocular nystagmus due to certain
positions of the head. Acta Otolaryngol (Stockh) 6: 106-137
Ogata Y, Sekitani T, Shimogori H, Ikeda T (1993) Bilateral vestibular neuronitis.Acta Otolaryngol (Stockh) SuppI503:57-60
Ohbayashi S, Oda M, Yamamoto M, Urano M, Harada K, Horikoshi
H, Orihara H, Kitsuda C (1993) Recovery of the vestibular function after vestibular neuronitis. Acta Otolaryngol (Stockh)
SuppI503:31-34
Ohm J (1932) ber die Beziehungen zwischen willkrlichen,
optischen und vestibulren Augenbewegungen. Z Hals Nas
Ohrenheilk 32: 234-246
Okinaka Y, Sekitani T, Okazaki H, Miura M, Tahara T (1993)

Vestibular neuritis
Progress of caloric response of vestibular neuronitis. Acta
Otolaryngol (Stockh) Suppl 503:18-22
Proctor L, Perlman H, Lindsay J, Matz G (1979) Acute vestibular
paralysis in herpes zoster oticus. Ann Otol Rhinol Laryngol
88:303-310
Ruttin B (1909) Zur Differentialdiagnose der Labyrinth- und
Hrnerverkrankungen. Z Ohrenheilk 57:327-331
Safran AB, Vibert D, Issoua D, Husler R (1994) Skew deviation
after vestibular neuritis. Am J OphthalmoII18:238-245
Sando I, Black FO, Hemenway WG (1972) Spatial distribution of
vestibular nerve in internal auditory canal. Ann Oto181:305
Schuknecht HF (1985) Neurolabyrinthitis. Viral infections of the
peripheral auditory and vestibular systems. In: Nomura Y (ed)
Hearing loss and dizziness, Igaku-Shoin, Tokyo, New York, pp
1-15
Schuknecht HF, Donovan ED (1986) The pathology of idiopathic
sudden sensorineural hearing loss. Arch Oto Rhino Laryngol
243:1-15
Schuknecht HF, Kitamura K (1981) Vestibular neuritis. Ann Otol
Rhinol Laryngol 90, Suppl 78: 1 - 19
Schuknecht HF, Will RL (1985) Acute bilateral sequential vestibular neuritis. Am J OtolaryngoI6:255-257
Schulz P, Arbusow V, Strupp M, Dieterich M, Rauch E, Brandt T
(1998) Highly variable distribution of HSV-l-specific DNA in
human geniculate, vestibular and spiral ganglia. Neurosci Lell
252: 139-142
Sekitani T, Imate Y, Noguchi T, Inokuma T (1993) Vestibular neuronitis: epidemological survey by questionnaire in Japan. Acta
Otolaryngol (Stockh) Suppl 503:9-12
Shimizu T, Sekitani T, Hirata T, Hara H (1993) Serum viral antibody titre in vestibular neuronitis. Acta Otolaryngol (Stockh)
Suppl 503:74-78
Shirabe S (1988) Vestibular neuronitis in childhood. Acta
Otolaryngol (Stockh) SuppI458:120-122
Silvoniemi P (1988) Vestibular neuronitis. An otoneurological
evaluation. Acta Otolaryngol (Stockh) SuppI453:1-72
Sioane PD Baloh RW, Honrubia V (1989) The vestibular system in
the elderly: clinical implications. Am J OtolaryngoI1O:442-449

81
Smith PF, Darlington CL (1994) Can vestibular compensation be
enhanced by drug treatment? J Vestib Res 4:169-179
Strupp M, Arbusow V, Dieterich M, Sautier W, Brandt T (1998a)
Perceptual and oculomotor effects of neck muscle vibration in
vestibular neuritis: Ipsilateral somatosensory substitution of
vestibular function. Brain 121: 677-685
Strupp M, Arbusow V, Maag KP, Gall C, Brandt T (1998b)
Vestibular exercises improve central vestibulospinal compensation after vestibular neuritis. Neurology 51:838-844
Strupp M, Jger L, Mller-Lisse U,ArbusowV, Reiser M, Brandt T
(1998c) High resolution Gd-DTPA MR imaging ofthe inner ear
in 60 patients with idiopathic vestibular neuritis: no evidence
for contrast enhancement of the labyrinth or vestibular nerve. J
Vestib Res 8:1-7
Suzuki J, Cohen B (1964) Head, eye, body and limb movements
from semicircular canal nerves. Exp NeuroI1O:333-405
Tahara T, Sekitani T, Imate Y, Kanesada K, Okami M (1993)
Vestibular neuronitis in children. Acta Otolaryngol (Stockh)
SuppI503:49-52
Tran Ba Huy P (1994) Physiopathology of peripheral nonMeniere's vestibular dis orders. Acta Otolaryngol (Stockh) Suppl
5l3:5-1O
Vibert D, Husler R, Safran AB, Koerner F (1996) Diplopia from
skew deviation in unilateral peripheral vestibular lesions. Acta
Otolaryngol (Stockh) 116:170-176
Wennmo C, Pyykk I (1982) Vestibular neuronitis. A clinical and
electro-oculographic analysis. Acta Otolaryngol (Stockh)
94:507-515
Yamanaka T, Sasa M,Amano T, Miyahara H, Matsunaga T (1995)
Role of glucocorticoid in vestibular compensation in relation to
activation of vestibular nucleus neurons. Acta Otolaryngol
(Stockh) Suppl 519:168-172
Zajtchuk J Matz G, Lindsay J (1972) Temporal bone pathology in
herpes oticus.Am Otol Rhinol LaryngoI81:331-338
Zee DS (1985) Perspectives on the pharmacotherapy of vertigo.
Arch OtolaryngoI3:609-612
Zee DS, Preiosi TJ, Proctor LR (1982) Bechterew's phenomenon in
a human patient. Ann Neuro112:495

Meniere's disease

Meniere first described this syndrome in 1861.lt is The clinical syndrome


characterised by fluctuating hearing loss, tinnitus,
and prolonged but gradually decreasing attacks of
vertigo and nystagmus, which may last several The classic triad is characterised by
hours. This striking triad of symptoms made
Meniere's disease the most popular and certainly
episodic vertigo,
over-diagnosed
vertigo
syndrome.
Mono fluctuating hearing loss, and
symptomatic cochlear or vestibular variants are also

tinnitus.
weil known. About 6% of the patients will develop
vestibular drop attacks, which generally abate spontaneously. Meniere's disease is the fourth most common cause of vertigo. lt chiefly affects those between Attacks
30 and 50 years of age. The frequency of attacks is
irregular, and there is a tendency to bilateral involve- More specifically, the typical attack is experienced as
me nt in about 30-60% of patients. Spontaneous
improvement can occur in a few years or even up to an initial sensation of fullness of the ear,
a reduction in hearing,
a decade after the onset of the condition.
The causative hydrops can result either from occurrence or increase of tinnitus
insufficient fluid resorption in the endolymphatic
sac or from a blockage of the endolymphatic duct. followed by
Its aetiology is still unknown, although scarring
labyrinthitis can lead to endolymphatic hydrops rotational vertigo,
(Schuknecht and Gulya 1983). Endolymphatic postural imbalance,
hydrops and periodic ruptures in the membranes nystagmus, and
separating endolymph from perilymph cause nausea after a few minutes.
endolymph discharge and intermediate potassium
palsy of vestibulocochlear nerve fibres (Dohlman Spontaneous nystagmus is always present during the
1976). A permanent fistula of the membranous attack. It is more commonly observed to beat away
labyrinth - spontaneous or surgical - permits grad- from the diseased ear than toward it. Clinical recordual release of the excessive amount of endolymph, ings from two patients whose nystagmus was docuthereby arresting the hydropic condition and mented near the very beginning of the vertiginous
Meniere's attacks (Schuknecht and Bartley 1985). attack initially showed an ipsiversive "irritative" type
Various medications and behavioural advice have of nystagmus, which reversed its direction to a conbeen recommended for management; the efficacy of traversive "paralytic" type within seconds to minutes
most (with the exception of betahistine and diuret- (Bance et a1. 1991). Reversal of the nystagmus in
ics) have remained unproven. The latest thinking on Meniere's attack is compatible with membranemanagement of rare cases involves intratympanic rupture-potassium "nerve palsy" (p. 89). A 3D analygentamicin or destructive surgical procedures. sis of nystagmus in three patients during acute
Differential diagnosis includes other recurrent Meniere's attacks revealed only horizontal and torvestibulopathies such as migrainous or vascular ver- sional components(Ohyama et a1. 1997), indicating
tigo attacks, perilymphatic fistulas, vestibular neuri- involvement of a11 three semicircular canal nerves
tis, vestibular paroxysmia, and familial episodic (vertical components of posterior and anterior semiataxia with vertigo.
circular canals cance! each other).
83

84

Single attacks generally have no prodromi or


recognisable preeipitating factors. They occur both
in daytime and at night, even du ring sleep. One-third
of the patients report an increased tinnitus and
hearing loss, as weIl as a subjective feeling of fullness
of the ear, which precedes the vertigo attack as a
kind of aura, similar to that in epilepsy. During the
vertigo attack, most patients with Meniere's disease
experience an increase in the severity of tinnitus and
hearing loss in comparison with the vertigo-free
interval. The rare, transient improvement of hearing
during the attack is known as the Lermoyez
phenomenon (Lermoyez 1919). Loss of conseiousness is not typical but may occur - Meniere originally
reported it in 1861 - as a secondary syncope together
with other vegetative manifestations (Pascher 1967).
The attacks slowly subside after a few hours, but
some dizziness and unsteadiness usually remain for
a few days.
The diagnosis of Meniere's disease becomes more
difficult when patients do not present with the classic triad of Huctuating hearing loss, tinnitus, and
episodic vertigo.
Monosymptomatic cochlear or vestibular manifestations are possible variants at the beginning of
the disease, depending on whether the hydrops predominantly affects the auditory or the vestibular
part of the labyrinth. Cochlear attacks may occur as
acute low-tone sensorineural hearing loss with spontaneous improvement; vestibular attacks may occur
as episodic vertigo. Assoeiation with aural fullness
or tinnitus will help to support the diagnosis.
Transtympanic electrocochleography has emerged
as a useful tool to assess endolymphatic hydrops in
"vestibular" Meniere's disease (Dornhoffer and
Kaufman-Arenberg 1993).

Auditory symptoms and signs in the vertigofree interval


In the early stages of the disease patients have no
symptoms in the vertigo-free interval. The majority,
however, subsequently develop slowly progressive
tinnitus and hearing loss, which Huctuate to an
unusual degree compared with other inner ear diseases. In the early stage hearing loss preferentially
involves the lower frequeneies; the higher frequeneies are affected later in the course of the disease
(Friberg et al. 1984; Thomas and Harrison 1971). The
"Hat type" of pure tone audiograms is most common, independently of the duration of Meniere's disease (Enander and Stahle 1967; Thomas and
Harrison 1971; Meyer zum Gottesberge and Stupp
1980). Exceptions are "rising curves" during the early
stages (Enander and Stahle 1967) or "peak audio-

Vertigo

grams" in later stages (Paparella et al. 1982). Acute


low-tone sensory hearing loss without vertigo may
be caused by endolymphatic hydrops, but clinically it
leads to cochlear or classic Meniere's disease only in
a minority of patients (Yamasoba et al. 1994). A retrospective study of 475 patients suffering from
Meniere's disease reported large variations of progressive hearing loss, but overall the low frequeneies
were more affected than the high ones, regardless of
the time elapsed since onset of the disease
(Katsarkas 1996). There is a positive recruitment
phenomenon with high sensitivity to loud sounds. A
positive Hennebert sign can be elicited in 30% of the
patients, because the fibrotic vestibulum allows
pathological transfer of pressure changes via the
tympanic membrane and the stares to the membranous labyrinth (Schuknecht 1976; NadoI1977). In
such cases, vertigo and nystagmus occur if negative
press ure is applied to the external auditory canal
(for example, by using a Politzer baIloon).
Klockhoff and LindbIom (1967) have shown that
when a transient hearing gain occurs after intake of
hyperosmolar substances (glycerol and/or urea)
endolymphatic hydrops is confirmed in about 60%
of the patients with Meniere's disease. Hearing losses
of more than 40 dB correlated with both a positive
glycerin test and a positive urea test (Imoto and
Stahle 1983). The standard criteria for positive
responses are either an improvement of 15 dB at a
frequency between 250 and 4000 Hz (pure tone +)
and/or a 12% improvement in speech discrimination
ability (speech +) (Snyder 1974). Onset and duration
of positive responses to the glycerin test vary
between 30 min and 3 h. Thus, the response can be
missed if the audiogram is done too soon or too late
(Lehrer and Poole 1982). Orally administered
glycerol does not always produce a significant
increase in plasma osmolality (Morrison et a1.l980).
Furthermore, it is not yet clear whether the osmotic
agent induces water transport between endolymphatic and perilymphatic space, or between vascular
and striate compartments (Duvall et al. 1980). The
observation that surgical withdrawal of endolymph
may worsen hearing, whereas glycerin improves it,
could indicate that the water displacement induced
by the osmotic test affects perilymph rather than
endolymph (Tran Ba Huy 1984). The acetazolamide
cochlear hydration test, which induces a worsening
of hearing, can help to detect latent hydrops
(Brookes et al. 1982).
The prognostic value of the glycerol dehydration
test and electrocochleography has been studied.
Abnormal electrocochleographic waveforms seem to
indicate a worse prognosis. After adequate dehydration, significant threshold changes for better or
worse predict the likely course of events following

85

Meniere's disease

endolymphatic sac surgery (Morrison 1986). Other


less unpleasant but still experimental procedures to
detect endolymphatic hydrops include measurement
of tympanic membrane displacements by the
stapedius reflex (Marchbanks 1984) or"low-frequency
masking" by applying a short acoustic stimulus and
a low frequency masker tone to the same ear in an
adjustable phase relationship (Mrowinski et al. 1996).

Vestibular function in the vertigo-free interval


The caloric responses of 475 patients in a retrospective study showed large variations, but overall they
deteriorated over time and did not correlate with the
level of hearing loss in most cases (Katsarkas 1996).
Vestibular function fluctuates - as does hearing loss
- with changes of endolymphatic pressure.
Furthermore, the furosemide- VOR test is positive in
more than 50% of patients with endolymphatic
hydrops (Ito et al. 1993). It has been shown in isolated
posterior semicircular canals of the frog that
ampullar receptors are inhibited by increasing
hydrostatic pressure due to the modified transmitter
release from the synaptic pole of the hair cells
(Zucca et al. 1991). Vestibular recruitment was not
demonstrated by sinusoidal rotational testing in
patients with Meniere's disease (Furman et al. 1990)
and was found in only 10-20% by using monothermal differential caloric testing (Wexler et al.
1991).

Imaging
Histological examination of the temporal bones
showed that patients with Meniere's disease were
more likely to have small vestibular aqueducts than
those without the disease (Sando and Ikeda 1984).
There are opposing views in the older literature
about the usefulness of visualising the aqueduct
(Hall et al. 1983 a,b) and about the significance of its
morphology. However, the advent of new visualising
techniques has supported the histological evidence.
For example, computed tomography of the petrous
bone has detected hypoplasia of the retrolabyrinthine region (Yazawa and Kitahara 1994), and
computed radiographic measurements of the
dimensions of the vestibular aqueduct in patients
with Meniere's disease have identified a hypoplastic
vestibular aqueduct with a narrow external aperture
(Takeda et al. 1997). Finally, high-resolution MR
imaging was able to visualise the endolymphatic
duct of patients with Meniere's disease significantly
less often than that of control subjects (Welling et al.
1996j Schmalbrock et al. 1996j Tanioka et al. 1997).

Differential diagnosis
There is no reliable clinical test to establish a diagnosis of Meniere's disease. It is easy to recognise if the
patient reports recurrent attacks with the typical
triad of symptoms (tinnitus, hearing loss, vertigo).
Frequently, however, endolymphatic hydrops must
be considered only one of many possible causes if a
monosymptomatic fluctuating hearing loss or vertigo
attacks occur without concomitant cochlear signs.
This and the varying severity and duration of the
attacks make the differential diagnosis of Meniere's
disease difficult.
Vertigo in migraine (benign paroxysmal vertigo
of childhood, basilar migraine, benign recurrent
vertigoj p. 325) must be considered as well as
vascular loop compression of the vestibular nerve
("vestibular paroxysmia"j p. 117) and idiopathic
recurrent vestibulopathy. Perilymph fistulas typically
manifest with combined auditory and vestibular
symptoms, and attacks can be induced by physical
exertion. Short Meniere's attacks (lasting a few minutes) must be distinguished from transient
ischaemia of the pontomedullary brainstem or the
labyrinth (vascular vertigoj p. 301) and from
vestibular paroxysmia (p. 117). Long Meniere's
attacks (lasting some days), especially if the attack is
the first one in the early stage of the disease, may be
indistinguishable from vestibular neuritis (p. 67)
including transient horizontal semicircular canal
paresis. If positional vertigo is prominent, benign
paroxysmal positioning vertigo must be excluded by
positioning manoeuvres. All kinds of infectious disorders of the inner ear, e.g. viral or bacterial
labyrinthitis/neuritis, can manifest with similar features of episodic vertigo and hearing loss, whereas
acoustic neurinoma (p. 155) usually does not produce vertigo attacks. Rare differential diagnoses
comprise familial episodic ataxia (p. 365), Cogan's
syndrome (p. 154), syphilitic labyrinthitis (p. 151),
vestibular atelectasis (p. 32), and hyperviscosity syndrome (p. 341).

Natural course
The incidence of Meniere's disease in a Swedish population was calculated to be 46/100 000, excluding
the cochlear type (with only fluctuating hearing
loss) (Stahle et al. 1978). Other reported incidence
rates vary from 21 to 50/100 000 (Dickins and
Graham 1990j Shojaku and Watanabe 1997). The preferred age of onset lies in the fourth to sixth decades,
and there is a slight preponderance of females

Vertigo

86

(Stahle et al. 1978). The significance of a statistical


predominance of the left ear in Meniere's disease,
sudden deafness, inner ear damage, tinnitus, and
abnormally patent Eustachian tube (Heermann
1993) remains obscure. The frequently positive family
history (Bernstein 1965; Birgerson et al. 1987) suggests genetic predisposing factors. Meniere's disease
occurs only rarely in children (Meniere 1861;
Beddoe 1977; Stahle et al. 1978; Parving 1976;
Meyerhoff et al. 1978; Watanabe 1981; Sade and Yaniv
1984). A large proportion of these children may be
labelled as suffering from "secondary or symptomatic" Meniere's syndrome (see delayed endolymphatic hydrops; p. 88) because of histories of an
initial hearing loss following mumps, Haemophilus
inJluenzae meningitis, temporal bone fractures, or
congenital or embryopathic complications in the ear
that developed into the full Meniere's triad 5-11
years later (Hausler et al. 1987).
In the course of the disease, tinnitus and fluctuating hearing loss sometimes announce the first vertigo attack years in advance, since the hydrops
normally begins in the pars inferior of the labyrinth
and the cochlear duct, with initial ruptures of
Reissner's membrane within the helicotrema. In this
case fullness of the ear is a characteristic sensation.
Purely vestibular attacks without cochlear symptoms are comparatively rare but may be the initial
manifestation at the beginning of the disease.
Meniere's disease usually begins in one ear with
an irregularly increasing and fortunately, after some
years, decreasing frequency of attacks. The major
reduction of hearing occurs within the first few years
of the disease (Enander and Stahle 1967; Hulshof
and Baarsma 1981; Friberg et al. 1984); the same is
true for the reduction in vestibular function as measured by caloric testing (Hulshof and Baarsma 1981).
The longer patients with Meniere's disease are followed, the greater is the percentage of those who
develop bilateral disease (Morrison 1986). In the earlier stages, when symptoms have been present for up
to 2 years, about 15% of the cases are bilateral, while
after one or two decades, 30-60% are bilateral
(Jongkees 1971; Stahle 1976a,b; Meyer zum
Gottesberge and Stupp 1980; Paparella and Griebie
1984; Morrison 1986; Kitahara 1991). The wide range
of incidence rates for bilateral Meniere's disease
results not only from various follow-up times but
also from the unavailability of reliable clinical criteria
for bilaterality. In a review of 67 temporal bone
autopsies with endolymphatic hydrops, 30% displayed bilateral involvement (Yazawa and Kitahara
1990). This finding stresses the importance of conservatism when considering surgical (destructive)
intervention for intractable disease in one ear. The
percentage of patients with this diagnosis who

underwent operations fell from about 20% in the


1960s to 11% in the 1970s (Stahle et al. 1978) and
declined further when regular "endolymphatic sac
operations" were found to have a placebo effect
(Thomson et al. 1981). The indication for surgical
intervention becomes even more doubtful (p. 92) if
one takes into account the relatively benign natural
his tory of the disease and its spontaneous remission
rate of about 80% within 5-10 years (Friberg et al.
1984). Remission may be achieved earlier with
endolymphatic-mastoid shunt surgery (Quaranta et
al. 1998). It is most likely that the spontaneous remittence of Meniere's attacks is due to permanent fistulisation between endo- and perilymph, allowing a
continuous asymptomatic leakage of excessive
endolymph.
A self-administered questionnaire to determine
dis ability in Meniere's disease revealed that the most
problematic symptom was vertigo, followed by hearing loss (Cohen et al. 1995). The unpredictability of
attacks and the lack of a safe place to rest during
attacks are significant problems for a few patients.
Psychological assessment supports the addition of
depression to the clinical picture of active Meniere's
disease (Coker et al. 1989).

Aetiology and pathomechanism


Endolymphatic hydrops
After the original description of the disease by
Prosper Meniere in 1861, the first step toward an
understanding of its pathology was the histological
finding of endolymphatic hydrops (Fig. 5.1) by
Hallpike and Cairns (1983) and Yamakawa (1938). In
a double-blind assessment of case histories and
histopathological findings in temporal bones, all 13
patients with clinical Meniere's syndrome were
found to have endolymphatic hydrops; however, the
medical records of six of 106 controls exhibited idiopathic endolymphatic hydrops without clinical
Meniere's syndrome (Rauch et al. 1989). Contrary to
earlier hypotheses, endolymphatic hydrops does not
develop as a consequence of increased endolymph
production or due to malfunction of the membrane,
which stabilises concentrations of electrolytes and
osmolarity between endolymph and perilymph
(Rauch 1968).
Impaired resorption of endolymph by the
endolymphatic sac is the major cause of hydrops.
This view is supported by electron-microscope
studies of the resorption mechanism (Lundquist
1976), temporal bone studies showing perisaccular

87

Meniere's disease
NORMAL

ENDOLYMPHATIC HYDROPS

Fig.S.l. Sehematie representation of the normal cochlea (Ieft) with typical histologie ehanges in endolymphatie hydrops (fight) 1,
endolymphatic hydrops, 2, 1055 of coehlear neurons; 3, atrophy of organ of Corti.

fibrosis (Hallpike and Cairns 1983; Altmann and


Zechner 1968; Zechner 1976), hypoplasia or atrophy
of the endolymphatic sac (Egami et al. 1978;
Arenberg et al. 1970), and the comparatively frequent inability of imaging techniques to visualise
the vestibular aqueduct (Welling et al. 1996; Tanioka
et al. 1997). Another pathogenic factor may be the
blockage of the longitudinal ftow of endolymph;
both Schuknecht (1977) and Zechner (1980) found
osseous blockage of the vestibular aqueduct.
Cochlear hydrops occurs with pathological obliteration of the saccule and/or the ductus reuniens
(Kitamura et al. 1982). This was observed in guineapigs after obliteration of the ductus reuniens
(Kimura et al. 1980) and after experimental perilymph fistulas (Nomura et al. 1987). Perilymph fistulas can be associated with hydrops (Kohut et al.
1986) (see perilymph fistula; p. 000). A purely apical
endolymphatic hydrops is of no pathological or
functional significance (Yamashita and Schuknecht
1982). Figures 5.2 and 5.3 show histological and
schematic morphological changes of the labyrinth at
different stages of the disease.
Naito (1950) was the first to produce an endolymphatic hydrops in guinea-pigs by ablation of the
endolymphatic sac. This was reproduced by Kimura
(1967) in the guinea pig as well as by Schuknecht et
al. (1968) in the cat, all of whom experimentally
blocked or obliterated the endolymphatic sac.
Surgically induced endolymphatic hydrops in the
guinea-pig causes cochlear and vestibular changes
similar to those observed in Meniere's disease, which
makes it a useful model according to Aran et al.
(1984). Andrews and Honrubia (1988), on the
contrary, point out that the guinea-pig model fails
to develop accompanying symptoms of the
attacks, such as sudden hearing loss or vestibular
imbalance.

Aetiology
Endolymphatic hydrops can be classified as embryopathic, acquired, or idiopathic, depending on the
aetiology (Schuknecht and Gulya 1983). The embryopathic type is rare and may be secondary to
Mondini dysplasia (Schuknecht 1980). In the
acquired type, a previous insult to the labyrinth can
usually be documented, either inftammatory (viral,
e.g. mumps; bacterial or spirochetal) or traumatic,
such as temporal bone fracture (Schuknecht and
Gulya 1983). Thus, viral neurolabyrinthitis may
result in endolymphatic hydrops (delayed endolymphatic hydrops from cochleovestibular labyrinthitis;
p.88), sensorineural hearing loss (cochlear
labyrinthitis), or episodic vertigo (vestibular neuritis)
(Schuknecht 1985). Inner ear autoimmunological
processes mayaiso be involved. Using the sensitive
(but unspecific) polyethylene glycol assay, Derebery
and colleagues (1991) found a significant elevation
of circulating immune complexes in patients with
Meniere's disease. Alleman et al. (1997) also detected
elevated autoantibodies to the endolymphatic sac in
such patients. If the fissure of a temporal bone fracture extends through the vestibular aqueduct, it can
produce fibro-osseous obliteration with secondary
impairment of endolymph resorption. The aetiology
of the idiopathic type of endolymphatic hydrops is
not known; the noxious factor does not cause an initially detectable disease with obvious cochlear or
vestibular dysfunction, and it is followed by
Meniere's syndrome.
Alternatively, however, endolymphatic hydrops
can be present but asymptomatic:
1. If it is not progressive, ruptures and distortion of

the membraneous labyrinth do not occur.


2. If no auditory or vestibular functions remain,

Vertigo

88

d
Fig.5.2. a Meniere's disease (mild hydrops, temporal bone section). There is a slight dilatation of the endolymphatic duct in the middie turn of the cochlea. H&E 42x b Meniere's disease (moderate hydrops, temporal bone section). A moderate dilatation of the
endolymphatic duct in the middle turn of the cochlea can be seen. H&E 12x. c Meniere's disease (severe hydrops, temporal bone section). The cochlear duct in the middle turn of this cochlea has dilated to such an extent that it has completely filled the scala vestibuli.
Reissner's membrane cannot be clearly seen because it is Iying against the roof of the scala vestibuli. H&E 42x. d Meniere's disease
(healed rupture, temporal bone section). Reissner's membrane has ruptured; however, the two margins of the rupture appear to have
healed together, leaving an invaginated stump. H&E 42x. (From Hawke and Jahn 1987.)

endolymphatic hydrops (constant cochlear and


vestibular defects) would not be symptomatic,
even if it were progressive.
3. If fistula formation has occurred spontaneously,
a sudden mixing of perilymph-endolymph is
prevented (Schuknecht 1976; Schuknecht and
Gulya 1983).
Delayed endolymphatic hydrops

Acquired types of endolymphatic hydrops are sometimes separated from idiopathic Meniere's disease
and called "delayed endolymphatic hydrops". Nadol
et al. (1975) first described this under a different
name as "vertigo of delayed onset after sudden deafness". Similarly, Wolfson and Leiberman (1975)
labelled it "unilateral deafness with subsequent vertigo". Schuknecht broadened this definition in 1978:
"it occurs in patients who have sustained a profound

hearing loss in one ear, usually from infection or


trauma, and then after a prolonged period of time
develop either episodic vertigo from the same ear
(ipsilateral delayed endolymphatic hydrops) or ftuctuating hearing loss, also sometimes with episodic
vertigo, in the opposite ear (contralateral delayed
endolymphatic hydrops)". This mechanism obviously
accounts for a number of Meniere's syndromes in
children (Hausler et al. 1987). The diagnosis is made
probable if the patient history is typical; the
furosemide test may be helpful to confirm the
hydropic condition (Futaki et al. 1984).
Vascular hypo thesis

The observation that migraine is often associated


with Meniere's disease and that treatment for
migraine mayaiso improve symptoms of Meniere's
disease (Parker 1995) is not a convincing argument

89

Meniere's disease

rI

./

fI~

11

Fig.5.3. a A diagram of the normal membranous labyrinth. b In the normal ear, so me hydrops has occurred. The ear can act swiftly
to remove the excess endolymph by increasing longitudinal flow. ein the Meniere's ear, a narrow endolymphatic duct silts up, preventing longitudinal flow and the endolymphatic hydrops increases. d The endolymphatic sac senses it is 'empty' and secretes glycoproteins and "saccin". e The obstruction is overcome with sudden onset of longitudinal drainage of endolymph towards the
endolymphatic sac. f Eventually the mechanisms fail and the duct remains blocked. There is gross endolymphatic hydrops but the
attacks of vertigo cease (burnt out Meniere's disease). (From Gibson and Arenberg 1997.)

for a common vascular pathophysiology. However,


its reported occurrence in one family allows the
speculation that Meniere's syndrome and migraine
have a common autosomal dominant genetic determinant (Oliveira et al. 1997).
Psychosomatic hypothesis

A number of publications advocate a psychosomatic


origin for Meniere's disease (Williamson and Gifford
1971; Elwood et al. 1982; Groen 1983). They base
their argument on case histories and the effects of
psychotherapy, but most of these publications suffer
from the absence of reliable scientific measurements
and conclusions (see also Psychogenic vertigo;
p. 453). Less positive reports on this topic clearly
state that the evidence supporting a psychosomatic
origin for Meniere's diseases is insufficient to justify
the conclusions drawn (Wexler and Crary 1986).

Pathophysiology of attacks and progressive


dysfunction
The literature contains controversial mechanical and
vascular explanations for the characteristic symp-

toms of hearing loss, tinnitus, and vertigo. Potassium


palsy produced by neurotoxic endolymph is another
disputed suggestion (Meyer zum Gottesberge and
Stupp 1980). Auditory signs and symptoms have all
been attributed to mechanical causes, namely alte red
cochlear-membrane macromechanics and/or ciliary
micromechanics (Tonndorf 1976,1983). The application of the mechanical hypotheses to explain
vestibular dysfunction as a simple consequence of
changes in inner ear pressure (Tonndorf 1983) cannot, however, explain the prolonged nystagmus and
vertigo typical of the attack. It is commonly accepted
that the latter are caused by periodic membrane ruptures that lead to transient potassium palsy of
vestibular nerve fibres. Vestibular organs and nerves
are located in the perilymph space, which contains
electrolyte concentrations similar to those of cerebrospinal fluid (Na+ = 143 mmolll, K+ = 8 mmol/l).
Potassium levels in the endolymph (K+ =
150 mmolll, Na+ = 15 mmolll) are high enough to
produce potassium-induced depolarisation ofaxons
causing conduction block (Smith et al. 1954).
Perfusion of these perilymphatic spaces with potassium solution can block cochlear responses (Tasaki
and Fernandez 1952) or cause paralytic nystagmus
(Dohlman 1965,1976; Silverstein 1970).

Vertigo

90

When potassium concentration increases, firstorder afferent nerve fibres passing through the perilymph are affected. Whereas they initially exhibit an
excitatory effect (increased spontaneous activity,
since the membrane potential is nearer to the activation potential of the sodium channel), if the concentration further increases, there is a blockade of the
action potentials (reduced spontaneous activity)
(Bance et al. 1991) due to continuous inactivation of
axonal sodium channels. This has been demonstrated
in guinea pigs by perfusion of the perilymph space
with artificial endolymph (Brown et al. 1988). Such a
mechanism explains why the nystagmus at the
beginning of the attack reverses from an initially
"ipsiversive irritative" to a subsequently "contraversive paralytic" type. It also explains the controversy
in the literature surrounding the direction of
nystagmus in the attack. The change in direction of
rotatory vertigo, postural imbalance, and nystagmus
may go undetected if the patient is not observed
within the first minute of the attack. There can be a
third (recovery) phase and a second revers al of the
nystagmus, which now beats again toward the affected
ear ("Erholungsnystagmus").
The direction of nystagmus and vertigo may in
addition depend on the location of the membranous
leakage in relation to either the posterior, anterior, or
horizontal ampullary nerve. Arecent 3D analysis of
spontaneous nystagmus in four patients with
Meniere's disease found, however, that there were
only two components of eye movements: horizontal
and torsional (Toshiaki et al. 1997). These findings
led Toshiaki and co-workers to speculate that all the
semicircular canal (horizontal and the two vertical)
afferents were involved in the attack. When both vertical canals are stimulated, the strong rotatory components prevail, since the counterdirected vertical
components cancel each other.
Continuous press ure on the sensory organs as
weIl as deformation of the labyrinth, including
spatial orientation of the otoliths (see vestibular
drop attacks; p. 94), by the hydrops can also cause
vestibular dysfunction. This dysfunction can either
fluctuate due to changes in hydrops pressure or manifest as a permanent vestibular defect. Schuknecht
(1984) has nicely formulated "a logical concept concerning the mechanism causing fluctuating hearing
loss and episodic vertigo of Meniere's disease:
1. Decreased endolymph resorption. Developmental

hypoplasia, trauma, or viral labyrinthitis cause


decreased resorption function of the endolymphatic sac.
2. Hydrops. There is a slow overaccumulation of
endolymph, causing hydrops and distortion of
the membranous labyrinth.

3. Ruptures. As the volume of endolymph increases,


there are repeated ruptures of the endolymphatic
system and contamination of the perilymphatic
fluid with neurotoxic endolymph, which temporarily paralyses the auditory and vestibular
mechanisms, causing vertigo andJor hearing loss.
4. Healing of the ruptures. The ruptures heal, which
allows the entire process to repeat itself.
5. Distortion and atrophy. With progression of the
disease, there are permanent alterations in the
biochemical and morphological features of the
membranous labyrinth which cause persistent
disequilibrium and/or hearing loss."

Management
The management of Meniere's disease has four aims:
1. to treat the acute attack,
2. to prevent further attacks,
3. to improve and/or preserve hearing and vestibular function, and
4. to prevent the development of bilateral Meniere's
disease.

To date, conservative and surgical procedures


have proved effective for only the first two aims.
There is much confusion in the literature about
which therapy is effective and which is the treatment
of first choice. No other vestibular dis order has been
the subject of such a large number of articles
(- 1500 between 1966 and 1996) which nevertheless
leave unanswered the questions of aetiology, pathomechanism, and effective treatment of the condition.
Patients should be given instructions on how to
manage the acute attack, and - most importantly they must be informed about the disease's overall
relatively benign course and in most cases the spontaneous remission or abatement of the attacks within a few years.

Attacks
The acute attack is self-limiting and subsides within
a few hours (rarely less than 1 h or more than a day)
in a slow decrescendo. The following recommendations can be made for management of the acute
attack:

With severe vertigo and postural imbalance,


patients should lie or sit down in order to prevent falls.

Meniere's disease

Head movements or rapid changes in head position should be restricted because of crosscoupled accelerations (Coriolis effects) and
positional vertigo.
If nausea is prominent, vestibular sedatives, such
as 50 mg dimenhydrinate (Dramamine), 4 mg
perphenazine (Fentazin), 25 mg promethazine
hydrochloride (Phenergan), or 0.6 mg scopolamine (Transderm-Scop), can be administered
parenterally (through the skin) for symptomatic
relief.

There is no justification for rheological infusions


or stellate ganglion blocks. Repeated caloric tests in
the attack-free interval have been proposed as
"deliberate dizziness therapy" for exceptional
patients in whom Meniere's vertigo attacks induce
panic reactions.

Attack-free interval
Treatment in the remission phase aims to reduce the
frequency of the attacks and to preserve hearing
without causing distressing tinnitus. Changing views
on the pathogenesis have prompted the development
of a variety of procedures. The existence of a large
nu~ber of different therapies, each defended fiercely
by its advocates, usually indicates that there is no
demonstrably effective therapy available. Dietetic
programmes including restriction of salt, water,
alcohol, nicotine, caffeine are as useless in treating
the disease as are physical exercise, avoidance of
exposure to low temperatures, or use of subatmospheric pressure chambers. Stellate ganglion
blocks, diuretics, vasoactive agents, tranquillisers,
neuroleptics, and lithium have been employed under
the questionable assumption that it is possible to
~iminish endolymphatic hydrops by changes in
mner ear blood flow, osmotic diuresis, or central
sedation. All these procedures (Stupp 1976; Pfaltz
1977; Schmidt 1977; Meyer zum Gottesberge and
Stupp 1980; Brandt and Bchele 1983) have been
fashionable therapies at one time or another, but with the possible exception of diuretics and betahistine - there has never been proof of their efficacy in
controlled prospective studies. In 1977, Torok surveyed 834 papers on the treatment of Meniere's disease which were published over a 25-year period and
concluded that "all have one common feature: all
claim success but not in 100%, recovery varies from
60 -80%." In 1991 Ruckenstein and colleagues
e~pressed a similar view: "This conclusion coupled
wlth the exposure of the lack of rationale behind
many of the proposed treatments, provided support
for the concept that patients with this disease benefitted

91

from a non-specific or placebo effect of therapy".


According to arecent review of the literature (Claes
and Van De Heyning 1997), only diuretics and
betahistine have a proven effect in double-blind
studies on the long-term control of vertigo, but no
medical therapy has a proven effect on hearing or
long-term evolution of the disease.
The effect of diuretics (acetazolamide) on experimental hydrops in guinea-pigs was confined to the
period of administration, and neither the administration nor non-administration of the drugs had any
effect on the extent of cochlear sensory and neural
atrophy in animals (Shinkawa and Kimura 1986).
Urea, thiazide diuretics, and acetazolamide have
been recommended for chronic treatment of
Meniere's disease. In a double-blind, placebocontrolled, crossover study, a combination of
tri amte rene and hydrochlorothiazide effectively
controlled vertigo (Van Deelen and Huizing 1986).
This combination, however, had no positive longterm effect on hearing, which confirms an earlier
report by Klockhoff et al. (1974) on long-term results
with chlorthalidone.
The histamine derivative betahistine (Aequamen,
Vasomotal) has been recommended as the drug of
first choice (Le Pere 1967; Chuden 1978). A one-year
prospective double-blind study concluded that this
treatment is preferable to no treatment (Meyer
1985). While it acts to improve the microcirculation
of the stria vascularis (Suga and Snow 1969;
Bertrand 1971) as a partial H 1 receptor agonist
(Laurikainen et al. 1993; Wang and Dutia 1995), it
also has inhibitory effects on polysynaptic vestibular
nucleus neurons (Unemoto et al. 1982). Other studies
support the view that betahistine is significantly
more effective than placebo (Oosterveld 1984), even
more effective than the diuretic hydrochlorothiazide
(Petermann and Mulch 1982) or flunarizine (Fraysse
et al. 1991). None of these studies document an
improvement in hearing loss. To prevent attacks,
betahistine should be administered continuously for
6-12 months.

Intratympanic gentamicin therapy


"Functionallabyrinthectomy" with ototoxic aminoglycosides (gentamicin or streptomycin), proposed
by Schuknecht in 1957, was first tried in Europe with
8-24 mg gentamicin sulphate (Refobacin) instilled
daily through a plastic tube inserted behind the anulus via the transmeatal approach (Lange 1977; Beck
and Schmidt 1978; Schmidt and Beck 1980). At that
time it was thought that it is possible to selectively
damage the dark cells of the secretory epithelium
(and thereby improve endolymphatic hydrops)

92

Vertigo

before significantly affecting vestibular and cochlear


function. Instillations were stopped when daily
audiograms or acheck of spontaneous nystagmus by
using Frenzel's glasses indicated a beginning endorgan deafferentation. Since then indications and
recommendations for intratympanic gentamicin
therapy (Graham and Goldsmith 1994; Hellstrm
and dkvist 1994; Halmagyi et al. 1994) have
changed, especially when Magnusson and Padoan
(1991) observed that the onset of ototoxic effects was
delayed by a few days to a week after gentamicin
instillation.
It is most likely that the route of transport of gentamicin from the middle to the inner ear is through
the round window membrane to the perilymphatic
space and from there to the hair cells in the
endolymphatic space (Bergenius and dkvist 1996).
Ototoxicity is probably caused by a reversible transduction channel blocking (Kroese et al. 1989) and by
damage to mitochondria due to excessive mitochondrial superoxide production that leads to cell death
(Hutchin and Cortopassi 1994). Protection from gentamicin ototoxicity by iron chelators was shown in
the guinea pig in vivo (Song et al. 1996; Song and
Schacht 1997). Intramuscular application of ototoxic
drugs will damage both labyrinths. Reports on therapy with streptomycin titration for bilateral
Meniere's disease (Langman et al. 1990) are not convincing.
These multis tage mechanisms of gentamicin ototoxicity are consistent with findings that the functional deficit was reversible at an early stage and
became irreversible at a late stage (Halmagyi et al.
1994). Application of excessive gentamicin can cause
unnecessary, inadvertent damage to the inner ear
receptors, including the cochlear hair cells. Low-dose
treatment - which does not even diminish or abolish
caloric responses of the treated ear - has also been
demonstrated to be effective (Yamazaki et al. 1991;
Murofushi et al. 1997; Driscoll et al. 1997) and is
therefore recommended as the standard procedure.
Rare indications for intratympanic gentamicin
therapy are the same as those for surgical
labyrinthectomy (Graham and Goldsmith 1994;
Bergenius and dkvist 1996; Murofushi et al. 1997):

conservatively or pharmacologically intractable


course, with frequent vertigo attacks or drop
attacks over more than 6 months, and hearing
loss to a non-serviceable level (hearing loss >60
dB) on the affected side;
continuing attacks despite selective vestibular
nerve section (a rare failure of vestibular
destructive procedures due to anatomic variants;
Monsei et al. 1988).

Since severe hearing loss does not always occur


when gentamicin is carefully instilled (dkvist 1988;
Nedzelski et al. 1993; Murofushi et al. 1997), some
otolaryngologists give ototoxic treatment even in
patients with moderate hearing loss, if hearing in the
opposite ear is unaffected. Bilateral manifestation of
Meniere's disease is a relative contraindication for
ototoxic treatment.
There is no general agreement on the optimal
concentration, temporal sequence, and total dosage
of intratympanal gentamicin instillations. Concentrations of 30 mg/mI gentamicin or less have usually
been administered (Magnusson and Padoan 1991;
Yamazaki et al. 1991; Nedzelski et al. 1993; Murofushi
et al. 1997). Two to three injections on consecutive
days were effective and had fewer side effects, such
as chronic hearing loss (dkvist 1988) or vestibular
insufficiency (Murofushi et al. 1997), than four or
more injections.
All the reported experience with this kind of
treatment indicates that one injection per week
(1-2 ml with concentrations less than 30 mg/mI) on
an outpatient basis could be recommended in order
to better monitor the delayed ototoxic effects.
Approximately 15% of patients with unilateral
vestibular deafferentation develop symptoms of
chronic vestibular insufficiency (Halmagyi 1994),
such as oscillopsia with head motion and unsteadiness during locomotion. This can be partly attributed to impaired vestibular function in the
remaining labyrinth of these patients and to incompIe te central compensation for loss of peripheral
function during rapid head movements (Aw et al.
1994).

Surgical treatments: non-destructive or


destructive
Non-destructive

Nowadays, fewer than 1-5% of patients ultimately


require surgical treatment, since the success of regular "endolymphatic sac shunt operations" (as modified after Portman 1927, and widely performed all
over the world for decades) was shown to be due to a
placebo effect (Thomson et al. 1981).
A double-blind, placebo-controlled study compared the efficacy of an endolymphatic sac-mastoid
shunt with a purely placebo operation (mastoidectomy)
in 30 patients with typical Meniere's disease. FoIlowup showed equally long-term improvement rates of
70% after 3 (Bretlau et al. 1984) and 9 years (Bretlau
et al. 1989). Despite all ongoing attempts to find the
optimal non-destructive endolymphatic shunt operation utilising valved shunts (Kaufmann-Arenberg et

Meniere's disease
al. 1988) or endolymphatic sac ballooning (Huang
and Lin 1994), a careful scrutiny of these procedures
has not shown either to be superior to placebo
(Ruckenstein et al. 1991; Schuknecht 1992). Although
single studies conclude that endolymphatic sac
shunt operations are effective as initial surgical procedure for long-term control of disabling vertigo
(Telischi and Luxford 1993), the opinion prevails even among surgeons - that the long-term results of
a shunt for relief of vertigo and preservation of hearing are not encouraging. The claim that long-term
bilaterality of Meniere's disease is less in surgical
versus non-surgical patients (Rosenberg et al. 1991)
does not agree with the literature.
Spontaneous permanent fistulisation is a possible
explanation for permanent recovery in Meniere's
disease. Thus, surgical fistualisation in various parts
of the membranous labyrinth has been used in
animal experiments (Kimura 1984) and in patients
with Meniere's disease. Cochlear endolymphatic
shunt operation has been tried, and Schuknecht and
Bartley (1985) report that over periods ranging from
1 month to 6 years (average 22 months; 102 ears)
72% of the cases were relieved of vertigo, but hearing
worsened in 45% of the cases.

Destructive
While surgical and other procedures that destroy the
peripherallabyrinth or vestibular nerve can successfully stop attacks of vertigo, they do not improve
hearing (Fisch 1976; House 1975). For patients in
whom all attempted conservative procedures have
failed, selectively destructive surgical techniques
(Van De Heyning et al. 1997), such as middle fossa
vestibular nerve section or ultrasonic or cryosurgical vestibular destruction, have been proposed to
preserve serviceable hearing function. As this surgical approach does not affect the hydrops pathomechanism and, therefore, does not prevent ongoing
fluctuating hearing loss, it is obviously not often
considered. The relatively benign natural history of
Meniere's disease should always be taken into
account and explained to the patient. Selective
chemical vestibulectomy was tried by placing a certain quantity of streptomycin between the bony and
the membranous part of the lateral semicircular
canal (Ecke et al. 1997).
Focused ultrasound seemed to have an advantage
over open surgery, since partial ablation of vestibular function (with preservation of hearing) can be
performed without invading the labyrinth. It has
been proposed as a useful treatment (Sjberg and
Stahle 1965; Angell-James 1970; Basek 1973; Stahle
1976b); however, Peron et al. (1983) were bothered
by the risk of facial palsy (via the lateral canal

93

approach) and doubted that selective ablation of the


vestibular sense organs could be achieved by current
techniques of ultrasound irradiation.
Finally, there is the rare patient who suffers from
unilateral Meniere's disease with frequent vertigo or
drop attacks and has no functional hearing on the
affected side. In such patients surgicallabyrinthectomy
is given careful consideration.
A circumspect weighing of the controversial
aspects of the current surgicalliterature allows two
simple statements:
There is no convincing indication for endolymphatic sac surgery in Meniere's disease.
2. f the various destructive procedures,
labyrinthectomy, cryosurgical techniques, and
ultrasonic irradiation have been almost abandoned in favour of the rare use of selective
vestibular nerve section in exceptional patients
with truly unmanageable disease.
1.

Retrosigmoid vestibular neurectomy seems to be the


most suitable technique for preserving hearing and
reducing postoperative morbidity (Silverstein et al.
1990; Glasscock et al. 1991; Kubo et al. 1995).
However, particularly in elderly patients, ablative
surgical procedures may cause long-lasting postural
imbalance, because of the reduced ability of central
mechanisms to compensate for the postoperative
vestibular tone imbalance.

Pragmatic therapy

Vestibular sedatives such as benzodiazepines,


dimenhydrinate or scopolamine effectively
reduce vertigo and nausea in the acute attack.
Betahistine is the drug of first choice for preventing vertigo attacks (8-16 mg/d for 6-12 months).
Diuretics are considered second choice for preventing vertigo attacks.
Combinations of betahistine and diuretics may
be tried if single-drug treatment fails.
Intratympanic gentamicin is the treatment of
first choice for preventing vertigo attacks or drop
attacks in rare patients with intractable and
frequent attacks lasting for more than 6-12
months and non-serviceable hearing in the
affected ear.
Selective vestibular neurectomy is considered
second choice in the same exceptional patients
or first choice in patients with moderate hearing
loss.

94

Vertigo

Vestibular drop attacks (Tumarkin's


otolithic crisis)

Table 5.1 summarises the information given in


this chapter about Meniere's disease.
Table 5.1.

Vestibular drop attacks can occur not only in the


later stages of endolymphatic hydrops (Turnarkin
1936; Kuh11980) but at any time during the course of
Meniere's disease (Jansen and Russel 1988). In
exceptional cases they may even be the initial manifestation (Baloh et al. 1990). Baloh and co-workers
found only 12 of 175 patients with Meniere's disease
over a lO-year period; Black et al. (1982) reported a
similar incidence (11 of 200 patients). The drop
attacks occur from a standing or sitting position
without typical triggers or prodromi. The patients
describe the typical features (Kohl 1980; Baloh et al.
1990) in the following ways:

they feIt they were being pushed or shoved to the


ground,or
the surroundings suddenly moved or tilted, causing their fall.

As distinct from syncopies or epileptic seizures,


there is no associated loss of consciousness, and
patients are able to stand up immediately. Contrary
to patients with transient upside-down vision or
room-tilt illusions (p.224), patients with drop
attacks fall without appropriate postural reflexes.
According to the pathophysiological viewpoint, sudden changes in endolymphatic fluid press ure cause
inappropriate end-organ stimulation that results in a
reflex-like vestibulospinalloss of postural tone or an
inappropriate vestibulospinal reflex that leads to a
fall.
In the series of Baloh and co-workers (1990), the
first drop attack occurred from less than 1 year to 29
years after onset of Meniere's disease, and the total
number of attacks varied from 2 to 18, with only 2 of
12 patients having more than six attacks. Drop
attacks tend to occur in a flurry during aperiod of 1
year or less and are followed by spontaneous remission (Jansen and Russel 1988; Baloh et al. 1990).
Therefore, conservative management is recommended,
not surgical intervention as proposed by Black et al.
(1982).
The pharmacological approach recommends
administering fentanyl and drop er idol (Innovar),
but this has had questionable success. Drop attacks
disappeared completely after gentamicin treatment
(intratympanally) (dkvist and Bergenius 1988).
Surgical treatment is the latest thinking here. It consists of either ipsilaterallabyrinthectomy or selective
section of the vestibular nerve in order to preserve
useful hearing in the affected ear (Black et al. 1982).

Meniere's disease

Clinical syndrome
- Fluctuating hearing 1055
- Tinnitus
- Subjeetive fullness of the ear
- Prolonged vertigo/nystagmus attacks with nausea
- Rare vestibular drop attaeks
Monosymptomatie forms possible, variable auditory and vestibular
deficits in the intervals between attaeks. There is no pathognomonie test
to establish the diagnosis unequivoeally.
Incidence/age/sex
- 50/100 000
- Affeets mainly age group from 30 to 50 years
- Incidenee in males and females roughly equal
- Rare in ehildren
Pathomechanism
- Endolymphatic hydrops of the labyrinth due to insufficient fluid
resorption in the endolymphatie sae or bloekage of longitudinal
endolymph flow
- Attacks. Periodie ruptures of the endolymph membrane with potassium
palsy of ampullary nerves and meehanieal hearing disturbance
-Intervals. Pressure-dependent 1055 of coehlear and vestibular neurons,
distortion of labyrinth struetures
Aetiology
- Aequired, "delayed endolymphatie hydrops"(i.e.labyrinthitis, viral or
baeterial; traumatic, temporal bone fraeture)
- Embryopathie (e.g. Mondini dysplasia)
- Idiopathie (aetiology not known)
Course/prognosls
- Usually beg ins in one ear with increasing frequeney of attaeks and major
auditory/vestibular deficit oecurring du ring the first years
- Thereafter spontaneous reduetion in vertigo attaeks (permanent
fistulisation ?), no further progression of deficit but inereasing
involvement of the opposite ear (30-60%)
Management
- Drugs
- Betahistine
- Diureties
- Destruetive (in rare eases)
- Ototoxie antibiotics (gentamicin)
- Vestibular nerve seetion
Differential diagnosis
- Vertigo in migraine (benign paroxysmal vertigo of ehildhood, basilar
migraine, benign reeurrent vertigo),
- Perilymph fistula,
- Neurovascular eompression ("vestibular paroxysmia"),
- Vestibular neuritis,
- Benign paroxysmal positioning vertigo,
- Transient isehaemie attaeks,
- Familial episodic ataxia,
- Cogan's syndrome,
- Syphilitie labyrinthitis,
- Vestibular ateleetasis,
- Hyperviscosity syndrome

Meniere's disease

References
Alleman AM, Dornhoffer JL, Kaufman Arenberg I, Walker PD
(1997) Demonstration of autoantibodies to the endolymphatic
sac in Meniere's disease. Laryngoscope 107:211-215
Altmann F, Zechner G (1968) The pathology and pathogenesis of
endolymphatic hydrops. New investigations. Arch Klin Exp
Ohr- Nas-Kehlk-Heilk 192:1-19
Andrews JC, Honrubia V (1988) Vestibular function in experimental endolymphatic hydrops. Laryngoscope 98:479-485
Angell-James J (1970) Erfahrungen bei der Behandlung der
Meniere'schen Krankheit mit Ultraschall. HNO 18:202-205
Aran J-M, Rarey KE, Hawkins JE (1984) Functional and morphological changes in experimental endolymphatic hydrops. Acta
Otolaryngol (Stockh) 97:547-557
Arenberg IK, Morowitz WF, Shambaugh GE Jr (1970) The role of
the endolymphatic sac in the pathogenesis of endolymphatic
hydrops in man. Otolaryngol (Suppl) 275:7-43
Aw ST, Halmagyi GM, Curthoys IS, Todd MI, Yavor RA (1994)
Unilateral vestibular deafferentation causes permanent impairment of the human vertical vestibulo-ocular reflex in the pitch
plane. Exp Brain Res 102:121-130
Baloh RW, Jacobson K, Winder T (1990) Drop attacks in Meniere's
syndrome. Ann Neurol 28:384-387
Bance M, Mai M, Tomlinson D, Rutka J (1991) The changing direction of nystagmus
in acute Meniere's
disease:
Pathophysiological implications. Laryngoscope 10 1: 197-201
Basek M (1973) Ultrasound for Meniere's disease. Lateral canal
versus round window approach. Arch Otolaryngol 97:133-134
Beck C, Schmidt CL (1978) Ten years of experience with intratympanally applied streptomycin (gentamicin) in the therapy of
morbus Meniere. Arch OtorhinolaryngoI221:149-152
Beddoe GM (1977) Vertigo in children. Otol Clin North Am
10:139-144
BergeniusI, dkvist LM (1996) Transtympanic aminoglycoside
treatment in Meniere's disease. In: Baloh RW, Halmagyi GM
(eds) Disorders of the vestibular system. Oxford University
Press, New York, pp 575-582
Bernstein J (1965) Occurrence of episodic vertigo and hearing
loss in families. Ann Otol Rhinol Laryngol 74: 101-111
Bertrand RA (1971) Modification of the vestibular function with
betahistine H Cl. Laryngoscope 81 :889-898
Birgerson L, Gustavson K-H, Stahle J (1987) Familial Meniere's
disease: a genetic investigation. Am J OtoI8:323-326
Black FO, Effron MZ, Burns DS (1982) Diagnosis in management
of drop attacks of vestibular origin: Tumarkin's otolithic crisis.
Otolaryngol Head Neck Surg 90:256-262
Brandt Th, Bchele W (1983) Augenbewegungsstrungen. Fischer,
Stuttgart, New York
Bretlau P, Thomsen J, Tos M, Johnsen NJ (1984) Placebo effect in
surgery for Meniere's disease: a three-year follow-up study of
patients in a double blind placebo controlIed study on
endolymphatic sac shunt surgery. Am J OtoI5:558-561
Bretlau P, Thomsen J, Tos M, Johnsen NJ (1989) Placebo effect in
surgery for Meniere's disease: nine year follow up. Am J Otol
10:259-261
Brookes GB, Morrison AW, Booth JB (1982) Acetazolamide in
Meniere's disease - evaluation of a new diagnostic test for
reversible endolymphatic hydrops. Otolaryngol Head Neck
Surg 90:358-366
Brown DH, McClure JA, Dowar-Zapolski Z (1988) The membrane
rupture theory of Meniere's disease - is it valid? Laryngoscope
98:599-601
Chden HG (1978) Erfahrungsbericht ber Betahistin Anwendung bei Morbus Meniere. Laryngol Rhinol 57:997-1007
Claes j, Van De Heyning PH (1997) Medical treatment of Meniere's

95
disease: a review ofliterature. Acta Otolaryngol (Stockh) Suppl
526:37-42
Cohen H, Ewell LR, Jenkins HA (1995) Disability in Meniere's disease. Arch Otolaryngol Head Neck Surg 121:29-33
Coker NJ, Coker RR, Jenkins HA, Vincent KR (1989) Psychological
profile of patients with Meniere's disease. Arch Otolaryngol
Head Neck Surg 115:1355-1357
Derebery Mj, Rao S, Siglock Tj, Linthicum FH, Nelson RA (1991)
Meniere's disease: an immune complex-mediated illness?
Laryngoscope 101 :225-229
Dickins JRE, Graham SS (1990) Meniere's disease 1983-1989. Am J
Otol11:51-65
Dohlman GF (1965) The mechanism of secretion and absorption
of endolymph in the vestibular apparatus. Acta Otolaryngol
59:275-288
Dohlman GF (1976) On the mechanism of the Meniere attack.
Arch Oto-Rhino-Laryngol212:301-307
Dornhoffer jL, Kaufman-Arenberg 1(1993) Diagnosis ofvestibular Meniere's disease with electrocochleography. Am J Otol
14:161-166
Driscoll CLW, Kasperbauer JL, Facer GW, Harner SG, Beatty CW
(1997) Low-dose intratympanic gentamicin and the treatment
of Meniere's disease: Preliminary results. Laryngoscope
107:83-89
Duvall A, Stanti PA, Hukee Mj (1980) Cochlear fluid balance. A
clinical research overview. Ann Otol (St Louis) 89:335
Ecke U, Begall K, Amedee RG, Norris CH, Mann WJ (1997)
Selective chemical vestibulectomy. ORL 59:209-214
Egami T, Sando I, Black FO (1978) Hypoplasia of the vestibular
aqueduct and endolymphatic sac in endolymphatic hydrops.
Trans Am Acad Ophthalmol OtolaryngoI86:327-339
Elwood S, Carlton J, Cliffe MJ (1982) A physiological contribution to
the management of Meniere's disease. Practitioner 226: 1149-1152
Enander A, Stahle j (1967) Hearing in Meniere's disease. Acta
Otolaryngol (Stockh) 64:543-556
Fisch U (1976) Die Chirurgische Behandlung des Morbus
Meniere. Arch Otorhinolaryngol212: 385-391
Fraysse B, Bebear jP, Dubreuil C, Berges C, Dauman R (1991)
Betahistine dihydrochloride versus flunarizine. A double-blind
study on recurrent vertigo with or without cochlear syndrome
typical of Meniere's disease. Acta Otolaryngol (Stockh) Suppl
490:1-10
Friberg U, Stahle j, Svedberg A (1984) The natural course of
Meniere's disease.Acta Otolaryngol (Stockh) SuppI406:72-77
Furman jMR, Durrant JD, Hyre R, Kamerer DB (1990) Vestibular
recruitment in Meniere's disease. Ann Otol Rhinol Laryngol
99:805-809
Futaki T, Yamane M, Kawabata I, Nomura Y (1984) Detection of
delayed endolymphatic hydrops by the furosemide test. Acta
Otolaryngol (Stockh) SuppI406:37-41
Gibson WPR, Arenberg IK (1997) Pathophysiologie theories in the
etiology of Meniere's disease. Otolaryngol Clin North Am
30:961-967.
Glasscock ME, Thedinger BA, Cueva RA, Jackson CG (1991) An
analysis of the retrolabyrinthine vs the retrosigmoid vestibular
nerve section. Otolaryngol Head Neck Surg 104:88-95
Graham MD, Goldsmith MM (1994) Labyrinthectomy. Indications
and surgical technique. Otolaryngol Clin North Am 27:325-335
Groen JJ (1983) Psychosomatic aspects of Meniere's disease. Acta
Otolaryngol (Stockh) 95:407-416
Hall SF, O'Connor AF, Thakkar CH, Wylie IG, Morrison AW
(1983a) Significance of tomography in Meniere's disease:
Visualization and morphology of the vestibular aqueduct.
Laryngoscope 93:1546-1549
Hall SF, O'Connor AF, Thakkar CH, Wylie IG, Morrison AW
(1983b) Significance of tomography in Meniere's disease: periaqueductal pneumatization. Laryngoscope 93: 1551-1553
Hallpike CS, Cairns H (1983) Observations of the pathology of
Meniere's syndrome. Proc R Soc Med 31:1317-1336

96
Halmagyi GM (l994) Vestibular insufficiency following unilateral
vestibular deafferentation. Aust J Otolaryngoll:510-512
Halmagyi GM, Fattore CM, Curthoys IS, Wade S (l994)
Gentamicin vestibulotoxicity. Otolaryngol Head Neck Surg
111:571-574
Hausler R, Toupet M, Guidetti G, Basseres F, Montandon P (l987)
Meniere's disease in children.Am J OtolaryngoI8:187-193
Hawke M, Jahn AF (1987) Diseases of the ear. Clinical and pathological aspects. Gower, New York, London
Heermann J (1993) Predominance ofleft ear in Meniere's disease,
sudden deafness, inner ear damage, tinnitus and abnormally
patent eustachian tube. Ear Nose Throat J 72:205-208
Hellstrm S, dkvist L (l994) Pharmacologic labyrinthectomy.
Otolaryngol Clin North Am 27:307-315
House W (l975) Meniere's disease: management and theory.
Otolaryngol Clin North Am 8:515-535
Huang TS, Lin CC (l994) Endolymphatic sac ballooning surgery
for Meniere's disease.Ann Otol Rhinol LaryngoI103:389-394
Hulshof JH, Baarsma EA (1981) Follow-up vestibular examination
in Meniere's disease. Acta Otolaryngol (Stockh) 91:397-401
Hutchin T, Cortopassi G (l994) Proposed molecular and cellular
mechanism for aminoglycoside ototoxicity. Antimicrob Agents
Chemother 38:2517-2520
Imoto T, Stahle J (1983) The clinical picture of Meniere's disease
in the light of glycerin and urea tests. Acta Otolaryngol
(Stockh) 95:247-256
Ito M, Watanabe Y, Shojaku H, Kobayashi H, Aso S, Mizukoshi K
(1993) Furosemide VOR test for the detection of endolymphatic
hydrops. Acta Otolaryngol (Stockh) SuppI504:55-57
Jansen VD, Russel RD (l988) Conservative management of
Tumarkin's otolithic crisis.J OtolaryngoI17:359-361
Jongkees LBW (l971) Some remarks on the patients suffering
from Meniere's disease. Trans Am Acad Opthalmol Otolaryngol
75:374-378
Katsarkas A (1996) Hearing loss and vestibular dysfunction in
Meniere's disease. Acta Otolaryngol (Stockh) 116: 185-188
Kaufmann-Arenberg I, Gibson WPR, Bohlen HKH (1988)
Improvements in audiometric and electrophysiologic parameters following non-destructive inner ear surgery utilising a
valved shunt for hydrops and Meniere's disease. In: Nadol JB Jr
(ed) Second International Symposium on Meniere's disease.
Kugler & Ghedini,Amsterdam, pp 545-561
Kimura R (l967) Experimental blockade of the endolymphatic
duct and sac and its effect on the inner ear of the guinea pig.
Ann Otol Rhinol Laryngol 76:664-687
Kimura RS (l984) Fistulae in the membranous labyrinth. Ann
Otol Rhinol Laryngol 93:36-43
Kimura RS, Schuknecht HF, Ota CY, Jones DD (1980) Obliteration
of the ductus reuniens. Acta Otolaryngol (Stockh) 89:295-309
Kitahara M (1991) Bilateral aspects of Meniere's disease. Acta
Otolaryngol (Stockh) SuppI485:74-77
Kitamura K, Schuknecht HF, Kimura RS (l982) Cochlear hydrops
in association with collapsed saccule and ductus reuniens. Ann
Otol Rhinol LaryngoI91:5-13
Klockhoff I, Lindbiom U (1967) Glycerol test in Meniere's disease.
Acta Otolaryngol SuppI224:449-451
Klockhoff I, Lindbiom U, Stahle J (1974) Diuretic treatment of
Meniere's disease: long term results with clorthalidone. Arch
Otolaryngoll00:262-265
Kohut RI, Hinojosa R, Budetti JA (1986) Perilymphatic fistula: A
histopathological study. Ann Otol Rhinol Laryngol 95:446-471
Kroese ABA, Das A, Hudspeth AJ (1989) Blockage of transduction
channels of hair cells in bullfrog's sacculus by aminoglycoside
antibiotics. Hear Res 37:203-218
Kubo T, Doi K, Koizuka I, Takeda N, Sugiyama N, Yamada K,
Kohmura E, Hayakawa T (1995) Assessment of auditory and
vestibular functions after vestibular neurectomy for Meniere's
disease.Acta Otolaryngol (Stockh) 115:149-153

Vertigo
Kuhl W (1980) Vestibular-cerebral syncopes. Dtsch Med
Wochenschr 105:41-42
Lange G (1977) Die intratympanale Behandlung des Morbus
Meniere mit ototoxischen Antibiotika. Laryng Rhinol
56:409-414
Langman AW, Kemink JL, Graham MD (1990) Titration streptomycin therapy for bilateral Meniere's disease. Follow-up report.
Ann Otol Rhinol Laryngol 99:923-926
Laurikainen EA, Miller JM, Ouirk WS, Kallinen J, Ren T, Nuttall
AL, Grenman R, Virolainen E (1993) Betahistine-induced vascular effects in the rat cochlea. Am J OtoI14:24-30
Le Pere DM (1967) Evaluation of a new symptomatic treatment
for Meniere's disease. Clin Med 74:63-64
Lehrer JF, Poole DC (1982) Onset and duration of positive
responses to the glycerin test in patients with Meniere's disease.
Am J OtolaryngoI3:262-263
Lermoyez M (1919) Le vertigo qui fait entendre. Presse Med 27:1
Lundquist P-G (1976) Aspects of endolymphatic sac morphology
and function.Arch Oto-Rhino-LaryngoI212:231-240
Magnusson M, Padoan S (1991) Delayed onset of ototoxic effects
of gentamicin in treatment of Meniere's disease. Acta
Otolaryngol (Stockh) 111:671-676
Marchbanks RJ (1984) Measurement of tympanic membrane displacement arising from aural cardiovascular activity, swallowing, and intra-aural muscle reflex. Acta Otolaryngol (Stockh)
98:119-129
Meniere P (1861) Memoire sur les lesions de I'oreille interne donnant lieu a des symptomes de congestion cerebrale apoplectiforme. Gaz Med Paris, Ser 3, 16:597-601
Meyer ED (1985) Zur Behandlung des Morbus Meniere mit
Betahistindimesilat (Aequamen) Doppelblindstudie gegen
Plazebo (crossover). Laryngol Rhinol OtoI64:269-272
Meyer zum Gottesberge A, Stupp H (1980) Meniere'sche
Krankheit. In: Zllner F (ed) Hals-Nasen-Ohrenheilkunde in
Praxis and Klinik Vo16, Ohr H. Thieme, Stuttgart, pp 38.1-38.31
MeyerhoffWL, Paparella MM, Shea D (1978) Meniere's disease in
children. Laryngoscope 88: 1504-1511
MonseIl EM, Brackmann DE, Linthicum FH Jr (1988) Why do
vestibular destructive procedures sometimes fai!? Otolaryngol
Head Neck Surg 99:472-479
Morrison AW (1986) Predictive tests for Meniere's disease. Am J
OtoI7:5-10
Morrison AW, Moffat DA, O'Connor AF (1980) Clinical usefulness
of electrocochleography in Meniere's disease: an analysis of
dehydrating agents. Otolaryngol Clin North Am 13:703-721
Mrowinski D, Scholz G, Krompass S, Nubel K (l996) Diagnosis of
endolymphatic hydrops by low-frequency masking. Audiol
Neurootoll:125-134
Murofushi T, Halmagyi GM, Yavor RA (1997) Intratympanic gentamicin in Meniere's disease: results of therapy. Am J Otol
18:52-57
Nadol JB (1977) Positive Hennebert's sign in Meniere's disease.
Arch Otolaryngoll03:524-530
Nadol JB Jr, Weiss AD, Parker SW (1975) Vertigo of delayed onset
after sudden deafness.Ann Otol Rhinol LaryngoI84:841-846
Naito T (1950) Experimental studies on Meniere's disease.
Otorhinolaryngol Soc Jpn 53:19-20
Nedzelski JM, Chiong CM, Fradet G, Schessel DA, Bryce GE,
pfeiderer AG (1993) Intratympanic gentamicin instillation as
treatment of unilateral Meniere's disease: update of an ongoing
study.Am J OtoI14:278-282
Nomura Y, Hara M, Funai H, Okuno T (1987) Endolymphatic
hydrops in perilymphatic fistula. Acta Otolaryngol (Stockh)
103:469-476
dkvist LM (1988) Middle ear ototoxic treatment for inner ear
disease.Acta Otolaryngol (Stockh) SuppI457:83-86
dkvist LM, Bergenius 0 (1988) Drop attacks in Meniere's disease.Acta Otolaryngol (Stockh) SuppI455:82-85

Meniere's disease
Ohyama Y, Yagi T, Ushio K, Suzuki K (1997) 3D analysis of nystagmus during peripheral vertiginous attaeks. Aeta Otolaryngol
(Stockh) Supp1528:77-79
Oliveira CA, Bezerra RL, Araujo MF, Almeida VF, Messias CI
(1997) Meniere's syndrome and migraine: ineidence in one
family. Ann Otol Rhinol Laryngol106:823-829
Oosterveld WJ (1984) Betahistine dihydrochloride in the treatment of vertigo of peripheral vestibular origin. A double-blind
plaeebo-eontrolled study. J Laryngol Oto198:37-41
Paparella MM, Griebie MS (1984) Bilaterality of Meniere's disease.
Acta Otolaryngol (Stockh) 97:333-337
Paparella MM, McDermott JC, de Sousa LCA (1982) Meniere's
disease and the peak audiogram. Arch Otolaryngoll08:555
Parker W (1995) Meniere's disease. Etiologic considerations. Arch
Otolaryngol Head Neck Surg 121:377-382
Parving A (1976) Meniere's disease in childhood. J Laryngol Otol
90:817-821
Pascher W (1967) Vestibulrer Anfallsschwindel und
Bewutseinsstrungen. Arch klin Exp Ohr-, Nas- Kehlk-Heilk
188:384-388
Peron DL, Kitamura K, Carniol PJ, Schuknecht HF (1983) Clinical
and experimental results with focused ultrasound.
Laryngoscope 93:1217-1221
Petermann W, Mulch G (1982) Zur Langzeittherapie des Morbus
Meniere. Betahistin-dihydrochlorid und Hydrochlorothiazid im
Wirkungsvergleich. Fortschr Med 100:431-435
Pfaltz CR (1977) Pathophysiologische Aspekte und therapeutische
Mglichkeiten beim Morbus Meniere. Laryng Rhinol
56:396-401
Portman G (1927) The saccus endolymphaticus and an operation
for draining the same for the relief of vertigo. J Laryngo142:809
Quaranta A, Marini F, Sallustio V (1998) Long-term outcome of
Meniere's disease: endolymphatic mastoid shunt versus natural
his tory. Audiol Neurootol 3:54-60
Rauch S (1968) Biochemical aspects of pathogenesis of Meniere's
disease. Otolaryngol Clin North Am 1:369-374
.
Rauch SD, Merchant SN, Thedinger BA (1989) Meniere's syndrome and endolymphatic hydrops. Double-blind temporal
bone study. Ann Otol Rhinol Laryngol 98:873-883
Rosenberg S, Silverstein H, Flanzer J, Wanamaker H (1991)
Bilateral Meniere's disease in surgical versus non-surgical
patients. Am J OtoI12:336-340
Ruckenstein MJ, Rutka JA, Hawke M (1991) The treatment of
Meniere's disease: Torok revisited. Laryngoscope 101:211-218
Sade J, Yaniv E (1984) Meniere's disease in infants. Acta
Otolaryngol (Stockh) 97:33-37
Sando I, Ikeda M (1984) The vestibular aqueduct in patients with
Meniere's disease. A temporal bone histopathological investigation. Acta Otolaryngol (Stockh) 97:558-570
Schmalbrock P, Dailiana T, Chakeres DW, Oehler MC, Welling DB,
Williams PM, Roth L (1996) Submillimeter-resolution MR of
the endolymphatic sac in healthy subjects and patients with
Meniere's disease.Am J NeuroradiolI7:1707-1716
Schmidt CL (1977) Aktuelle medikamentse Therapie beim
Morbus Meniere. Laryng Rhino156:407-409
Schmidt CL, Beck CHL (1980) Behandlung des Morbus Meniere
mit intratympanal appliziertem Gentamycin-Sulfat. Laryng
Rhinol 59:804-807
Schuknecht HF (1957) Ablation therapy in the management of
Meniere's disease.Acta Otolaryngol (Stockh) Suppl132:1
Sehuknecht HF (1976) Pathophysiology of endolymphatic
hydrops. Arch Oto-Rhino-Laryngol212:253-262
Schuknecht HF (1977) Pathology of Meniere's disease as it relates
to the sac and tack proeedures. Ann Otol Rhinol Laryngol
86:677-682
Schuknecht HF (1978) Delayed endolymphatic hydrops. Ann Otol
Rhinol LaryngoI87:743-748
Sehuknecht HF (1980) Mondini dysplasia. A clinical and pathological study. Ann Otol Rhinol Laryngol Suppl 65:89

97
Schuknecht HF (1984) The pathophysiology of Meniere's disease.
Am J Otol 5:526-527
Schuknecht HF (1985) Neurolabyrinthitis. Viral infections of the
peripheral auditory and vestibular systems. In: Nomura Y (ed)
Hearing loss and dizziness. Igaku -Shoin, Tokyo, pp 1-15
Schuknecht HF (1992) Myths in neurootology. Am J Otol
13:124-126
Schuknecht HF, Gulya AJ (1983) Endolymphatic hydrops. An
overview and classification.Ann Otol Rhinol Laryngo192:1-20
Schuknecht HF, Bartley M (1985) Cochlear endolymphatic shunt
for Meniere's disease. J Otol (Suppl): 20-22
Schukneeht HF, Northrop C, Igarashi M (1968) Cochlear pathology
after destruetion of the endolymphatic sac in the cat. Acta
Otolaryngo165:479-487
Shinkawa H, Kimura RS (1986) Effect of diuretics on endolymphatic hydrops. Acta Otolaryngol (Stockh) 101:43-52
Shojaku H, Watanabe Y (1997) The prevalence of definite cases of
Meniere's disease in the Hida and Nishikubiki districts of central Japan: A survey of relatively isolated areas of medical care.
Acta Otolaryngol (Stockh) Supp1528:94-96
Silverstein H (1970) The effects of perfusing the perilymphatic
space with artificial endolymph. Ann Otol Rhinol Laryngol
79:754-765
Silverstein H, Norrel H, Rosenberg S (1990) The resurrection of
vestibular neurectomy: a lO-year experience with 115 cases. J
Neurosurg 72:533-539
Sjberg A, Stahle J (1965) Treatment of Meniere's disease with
ultrasound. Arch Otolaryngo182:498-502
Smith CA, Lowry OH, Wu M-L (1954) The electrolytes of the
labyrinthine fluids. Laryngoscope 64: 141-153
Snyder J (1974) Extensive use of a diagnostic test for Meniere's
disease. Arch Otolaryngol100:360-365
Song BB, Schacht J (1996) Variable efficacy of radical seavengers
and iron chelators to attenuate gentamicin ototoxicity. Hear Res
94:87-93
Song BB, Anderson DJ, Schacht J (1997) Protection from gentamicin ototoxicity by iron chelators in guinea pig in vivo. J
Pharmacol Exp Ther 282:369-377
Stahle J (1976a) Advanced Meniere's disease. A study of 356
severely disabled patients. Acta Otolaryngol (Stockh)
81:113-119
Stahle J (1976b) Ultra sound treatment of Meniere's disease. Longterm follow-up of 356 advanced cases. Acta Otolaryngol
(Stockh) 81:120-126
Stahle J, Stahle Ch, Arenberg IK (1978) Incidenee of Meniere's disease. Arch Otolaryngoll 04:99-102
Stupp H (1976) Die medikamentose Therapie der Meniere'schen
Krankheit. Arch Oto-RhinolaryngoI212:375-384
Suga J, Snow JB (1969) Cochlear blood flow in response to vasodilating drugs and some related agents. Laryngoscope
79:1956-1979
Takeda T, Sawada S, Kakigi A, Saito H (1997) Computed radiographie measurement of the dimensions of the vestibular
aqueduct in Meniere's disease. Acta Otolaryngol (Stockh) Suppl
528:80-84
Tanioka H, Kaga K, Zusho H, Araki T, Sasaki Y (1997) MR of the
endolymphatic duct and sac: Findings in Meniere's disease. Am
J NeuroradioI18:45-51
Tasaki I, Fernandez C (1952) Modification of cochlear microphonics and action potentials by KCL solution and by direct
currents. J NeurophysiolI5:497-512
Telischi FF, Luxford WM (1993) Long-term effieacy of endolymphatic sac surgery for vertigo in Meniere's disease. Otolaryngol
Head Neck Surg 109:83-87
Thomas K, Harrison MS (1971) Long-term follow-up of 610 cases
of Meniere's disease. Proc R Soc Med 64:853
Thomson J, Bretlau P, Tos M, Johnson NJ (1981) Placebo effect in
surgery for Meniere's disease. Arch Otolaryngoll07:271-277

98
Tonndorf J (1976) Endolymphatic hydrops: mechanical causes of
hearing loss. Arch Oto-Rhino Laryngol212:293-299
Tonndorf J (1983) Vestibular signs and symptoms in Meniere's
disorder: Mechanical considerations. Acta Otolaryngol
(Stockh) 95:421-430
Torok N (1977) Old and new in Meniere's disease. Laryngoscope
87:1870-1877
Toshiaki Y, Yoshio 0, Kayo S, Eriko K, Takayuki K (1997) 3D
analysis of nystagmus in peripheral vertigo. Acta Otolaryngol
(Stockh) 117:135-138
Tran Ba Huy P (1984) Electrophysiological and biochemical findings in four cases of Meniere's disease. Acta Otolaryngol
(Stockh) 97:571-579
Tumarkin A (1936) The otolithic catastrophe: a new syndrome. Br
Med J 1:175-177
Unemoto H, Sasa M, Takaori S, Ito J, Matsuoka 1 (1982) Inhibitory
effect of betahistine on polysynaptic neurons in the lateral
vestibular nucleus. Arch Otolaryngol 236:229-236
Van de Heyning PH, Verlooy 1, Schatte man I, Wuyts FL (1997)
Selective vestibular neurectomy in Meniere's disease: A review.
Acta Otolaryngol (Stockh) SuppI526:58-66
Van Deelen GW, Huizing EH (1986) Use of a diuretic (dyazide) in
the treatment of Meniere's disease. A double-blind cross-over
placebo-controlled study. ORL J Otorhinolaryngol Relat Spec
48:287-292
Wang n, Dutia MB (1995) Effects of histamine and betahistine on
rat medial vestibular nucleus neuron es: possible mechanism of
action of anti-histaminergic drugs in vertigo and motion sickness. Exp Brain Res 105:18-24
Watanabe 1 (1981) Meniere's disease in males and females. Acta
Otolaryngol (Stockh) 91:511-514
Welling DB, Clarkson MW, Miles BA, Schmalbrock P, Williams
PM, Chakeres DW, Oehler MC (1996) Submillimeter magnetic
resonance imaging of the temporal bone in Meniere's disease.
Laryngoscope 106: 1359-1364
Wexler M, Crary WG (1986) Meniere's Disease: The psychosomatic
hypothesis. Am J Otol 7:93-96

Vertigo
Wexler DB, Harker LA, Voots RJ, McCabe BF (1991) Monothermal
differential caloric testing in patients with Meniere's disease.
Laryngoscope 101:50-55
Williamson DG, Gifford F (1971) Psychosomatic aspects of
Meniere's disease.Acta Otolaryngol (Stockh) 72:118-120
Wolfson R1, Leibermann A (1975) Unilateral deafness with subsequent vertigo. Laryngoscope 85: 1762-1766
Yamakawa K (1938) ber die pathologische Vernderung bei
einem Meniere-Kranken. J Otorhinolaryngol Soc Jpn
4:2310-2312
Yamashita T, Schuknecht HF (1982) Apical endolymphatic
hydrops. Arch OtolaryngoI108:463-466
Yamasoba T, Kikuchi S, Sugasawa M, Yagi M, Harada T (1994)
Acute low-tone sensorineural hearing loss without vertigo.
Arch Otolaryngol Head Neck Surg 120:532-535
Yamazaki T, Hayashi M, Komatsuzaki A (1991) Intratympanic
gentamicin therapy for Meniere's disease placed by a tubal
catheter with systematic isosorbide. Acta Otolaryngol (Stockh)
SuppI481:613-616
Yazawa Y, Kitahara M (1990) Bilateral endolymphatic hydrops in
Meniere's disease: review of temporal bone autopsies. Ann Otol
Rhinol Laryngol 99:524-528
Yazawa Y, Kitahara M (1994) Computerized tomography of the
petrous bone in Meniere's disease. Acta Otolaryngol (Stockh)
SuppI51O:67-72
Zechner G (1976) Pathohistologie des Ductus and Saccus
endolymphaticus
beim
Innenohrhydrops.
Arch
Otorhinolaryngol 212:277-286
Zechner G (1980) Innenohrhydrops als Folge gestrter
Endolymphzirkulation. Laryng Rhinol 59:829-833
Zucca G, Botta L, Mira E, Manfrin M, Poletti A, Buizza A, Valli P
(1991) Effects of hydrostatic pressure on sensory dis charge in
frog semicircular canals. Acta Otolaryngol (Stockh)
111:820-826

Perilymph fistulas (PLF)

The perilymph space surrounds the endolymphfilled membranous labyrinth, and both are encapsuled by the bony labyrinth. Perilymph fistulas (PLF)
- abnormal communieations between the perilymph
space and the middle ear (Fig. 6.1) - are caused by
traumatic pressure changes in either the cerebrospinal fluid (explosive force) and/or the middle
ear (implosive force) (Fig. 23.1; p. 353). PLF may lead
to episodic vertigo and sensorineural hearing loss,
owing to pathologieal elasticity of the otie capsule or
leakage of perilymph, usually at the oval and round
windows. The fistula and a partial collapse of the
membranous labyrinth ("floating" labyrinth) permit
abnormal transfer of ambient pressure changes to
maculae and cupulae receptors.
The typical history is that of an "otolithic ataxia",
or a semicircular canal type of vertigo, and/or a
sudden hearing loss resulting from barotrauma (flying, diving; p. 351), trauma to the head, to the ear
(e.g. post-surgery), or from strenuous activity, such
as lifting of heavy weights (excessive Valsalva
manoeuvre). As trauma is a frequent aetiology of the
first manifestations of PLF, the subsequently vulnerable patients often report on typical triggers (lifting
weights, nose blowing, travelling through mountains) that set off the clinieal signs of episodic vertigo and/or sensorineural hearing loss. In some
patients PLF appear as sound-induced vestibular
symptoms, which are called the Tullio phenomenon
(p. 106), either of the semicircular canal or otolith type.
PLF probably account for a considerable proportion of those patients presenting with vertigo of
unknown aetiology, partieularly in children with
episodic vertigo and sensorineural hearing loss. This
is because of the great variability of signs and symptoms and the lack of a pathognomonie test. Surgical
exploration by tympanotomy is necessary in order
to establish the diagnosis. CT and MRI sometimes
reveal causative inner or middle ear abnormalities
or air bubbles. Detection of the specific CSF and
perilymph protein beta-2-transferrin in the middle
ear suggests perilymph leakage.
In the acute case, conservative treatment is
99

universally recommended since these fistulas usually


heal spontaneously. The prognosis for vertigo is
good, and patients with only mild hearing 10ss in the
symptomatic phase generally recover their hearing.
In contrast, the results of surgical interventions are
not encouraging as regards both the recurrence rate
and further improvement of the natural (spontaneous) course.

Round Window Membrane


Fig. 6.1. Membranous and bony labyrinths, vertiea l seetion,
human, right. This view shows the anatomie rela tionsh ip of parts
of the membranous labyrinth to the oval and round windows and
the middle ear. (Redrawn from Sehukneeht 1993.)

The clinical syndromes


The clinical picture of PLF is characterised by a wide
range of symptoms: pure vestibular symptoms; pure
hearing loss; combinations of both, including tinnitus and fullness of the ear; or the absence of symptoms. Patient history is very important, especially if
the first manifestation is associated with head
trauma.
With respect to vertigo and vestibular function
two types of PLF can be distinguished:

100
1. the semicircular canal type by rotational vertigo

and nystagmus,
2. the otolith type by unsteadiness, gait ataxia, and
oscillopsia.
Both types manifest in episodes lasting from hours
to days. Frequent triggers are ambient pressure
changes transferred to the inner ear, certain head positions in space, head movements, or locomotion.

Semicircular canal type of PLF


Vestibular dysfunctions are far more frequent than
hearing loss (Thompson and Kohut 1979; Singleton
et al. 1978; Simmons 1982; Anon and Miller 1985). In
the semicircular canal type of PLF, they are
described as episodic rotational vertigo with spontaneous nystagmus. These dysfunctions are sometimes precipitated by changes in "barotraumatic"
pressure (e.g. when travelling through mountains) or
position, such as bending over or assuming a supine
position with lateral head tilt. Rotational plane,
direction of apparent rotation, and nystagmus
depend on the site of the fistula relative to the
ampulla of the canal. This type of PLF may be confused with a short Meniere attack or neurovascular
compression (vestibular paroxysmia; p. 117). Positional
nystagmus is often observed and must be differentiated from benign paroxysmal positional vertigo
(BPPV) (Goodhill et al. 1973; Healy et al. 1974). In
PLF the short latency and long duration of positional
nystagmus, which is less violent than in BPPV
(Singleton et al. 1978), is not always present in our
experience.

Otolith type of PLF


Healy and co-workers (Healy et al. 1973; 1974) were
the first to stress that this condition should be suspected in patients with severe gait disturbance and
ataxia without evidence of central nervous system
disease, even in the absence of gross hearing defects.
In our experience, these patients represent a welldefined subgroup of fistula patients. It is justified to
call this syndrome "otolith type of PLF", since the
vertigo symptoms can be explained by inadequate
otolithic stimulation secondary to oval window fistulas. Most of these patients are free from vertigo with
the head stationary, but they experience a distressing
to-and-fro movement of both body and surroundings with head accelerations, for example, when rising from a sitting position and particularly when
walking. The sensation, described as "walking on
pillows", is similar to that described by patients in

Vertigo
Table 6.1. Perilymph fistulas (PLF)
Cliniml syndromes
A variety of episodic rotary or linear vertigo (often positional) associated
with sensorineural hearing loss, tinnitus, and ear pressure, especially
following excessive physical activity, head trauma, or barotrauma
Semicircular canal type
Episodic rotatory vertigo and nystagmus frequently associated with
sensorineural hearing loss, tinnitus, and ear pressure
Otolith type
- To-and-fro vertigo, unsteadiness, gait ataxia
- Oscillopsia with linear head motion
- Episodic vertigo is frequently induced by strenuous exercise,
lifting, diving, flying, bouts of sneezes, coughing
- Distressing vertigo and unsteadiness are frequently modulated by
head motion or head position in space
Diagnostie aids
History of head, ear, or barotrauma
Pressure fistula tests
Vascular fistula tests
Cl and MR imaging
Hyporesponsiveness to caloric irrigation
Increased postural sway with high-intensity sound stimulation
Electrocochleography
Audiological testing
Most reliable are
Exploratory tympanotomy
Detection of beta-2-transferrin in middle ear
Incidence/age/sex
Incidence and prevalence not known, probably more frequent than suspected
Occurrence throughout life with a relevant peak in childhood with no sex
preference
Pathomechanism
Pathological elasticity of the otie capsule and leakage of perilymph due to
rupture of oval, round, or both windows by "implosive" or "explosive" forces
(increased cerebrospinal fluid pressure)
The fistuled, partially collapsed labyrinth causes inadequate semicircular
canal and otolith stimulation, particularly with head motion and ambient
pressure changes ("floating" labyrinth).
Aetiology
Mostly traumatic, such as strenuous physical activity (Valsalva manoeuvre,
lifting heavy objects, bouts of sneezing), barotrauma (flying, diving), head
or ear trauma, surgical (stapedectomy, mastoideetomy, coehlear implants),
inflammatory (cholesteatoma, chronic otitis media), congenital (children)
Course/prognosis
High probability of spontaneous healing with resolving vertigo,
disequilibrium, and also sensorineural hearing loss
Management
1. Conservative treatment: bedrest with head elevation, avoidance of
straining, sneezing, coughing, and the use of stool softeners
2. Surgical patching offistula with subsequent medieal treatment for
stabilisation of healing
Treatment of first choice is conservative rather than surgical
Differential diagnosis
- Postconcussional syndrome
- Benign paroxysmal positioning vertigo
- Vestibular paroxysmia
- Meniere's disease
- Bilateral vestibulopathy
- Vestibular atelectasis
- Phobie postural vertigo
Acoustic vertigo: Tullio phenomenon
Sound-induced vestibular symptoms, such as vertigo, nystagmus,
oscillopsia, and postural imbalance in perilymph fistulas

101

Perilymph fistulas

the initial phase of BPPV (p. 256) or in phobic postural vertigo (p. 469). The gait is broad-based and
ataxic, but clinical examination does not reveal cerebellar or spinal ataxia. It is striking that head movements, which preferentially stimulate the canals
(horizontal oscillation in yaw), are much better tolerated than linear accelerations. Sometimes a linear
vertigo, described as a tilt or slow falling, is precipitated in the supine position (especially with the
affected ear undermost). Nausea and vomiting are
rare, unlike in canal disease. Symptoms most often
associated with this otolithic vertigo are ftuctuating
fullness of the ear, tinnitus, and sensorineural hearing loss.
The disease is more often episodic than chronic.
Episodes are sometimes induced by strenuous activities such as lifting heavy objects, jogging, or all
kinds of Valsalva press ure increases (sneezing,
coughing). The severity of the episodes varies. Some
patients, who are able to detect the beginning of an
episode by an audible "pop" or increasing fullness of
the ear, can prevent the development of more severe
symptoms merely by stopping the precipitating
activity.

How maya perilymph fistula be identified?

causes an ampullopetal shift of the endolymph, with


the nystagmus beating toward the pressurised ear. A
more sensitive fistula test has been developed by
using posturography with partial removal of both
visual and support surface orientation references of
the subject. This gives vestibular control a greater
sensorial weight in postural sway regulation (Black
et al. 1987). Most patients who subsequently underwent operation for PLF exhibited increased postural
sway with sinusuoidal changes in external auditory
canal pressure, but the specificity of the test could
not be determined because surgical tympanotomy
was not performed in all patients. Another modification of pressure fistula tests has been tried which
uses an impedance bridge (sudden pressure changes
from + 200 mm HzO to -400 mm H 20) and electronystagmographical recordings (Daspit et al. 1980).
Ocular torsion was recorded with a magnetic scleral
search coil by applying positive air pressure to one
extern al canal (Ostrowski et al. 1997). The superior
pole of the eye rotated away from the stimulated ear.

Vascular fistula tests


Vascular fistula tests (Mygind 1918; Nylen 1923) are
positive when bilateral compression of the jugular
veins causes eye movements or vertigo (Fig. 6.2)

It is unfortunately typical for medical diagnostics

that the number of proposed clinical tests and signs


for a particular disorder is inversely proportional to
their sensitivity and selectivity. We consider careful
his tory taking to be most important. We do not
advocate exploratory tympanotomy in uncomplicated
cases, since conservative management and observation of the mostly benign spontaneous course
should be preferred to surgical intervention. The following signs and tests may be helpful to support a
diagnosis of suspected PLF.
Press ure fistula tests

Pressure fistula tests (Lucae 1881), also known as the


Hennebert sign (Hennebert 1905), are positive when
pressure changes within the extern al auditory canal
evoke ocular deviation, nystagmus, oscillopsia, vertigo, or postural imbalance. This test is performed
either with Frenzel's glasses and a simple Politzer
balloon (Fig. 6.2) or with a more sophisticated pneumatic otoscope (Moon and Hahn 1978), which gives
positive results in only about 25% of cases (Singleton
et al. 1978; Simmons 1982). The direction of eye
movements or nystagmus depends on the location of
the fistula relative to cupula or macula. A fistula sign
is said to be typical (occur most frequently), when a
pressure increase in the external auditory canal

Fig.6.2. Schematic representation of typical pressure and vascular fistula signs in patients with perilymph fistulas. Pressure
changes in the external auditory canal (produced by a Politzer
balloon or a pneumatic otoscope) can stimulate the cupula of the
semicircular canal and/or the otoliths directly through either a
pathologically elastic bony labyrinth or through leakage (pressure fistula test). Similar effects are observed in response to bilateral compression of the jugular veins (increasing the intracranial
pressure) or with the head hanging down (vascular fistula test,c).

Vertigo

102

secondary to an increase of intracranial pressure


increments, which are pathologically transmitted
through the fistula. Valsalva manoeuvres can give
positive findings by a similar mechanism (Borries
1923) as can positional manoeuvres with the head
hanging down (Fig. 6.2; Stenger 1953). The diagnostic value of all fistula tests, however, is limited, since
a negative result does not exclude fistulas, and falsepositive results have been described as pseudo-fistula
signs in otosclerosis, Meniere's disease, syphilis, and
middle ear infections with disturbance of ventilation
through the Eustachian tube.
Imaging techniques (Cr, MRI)

High-resolution computed tomography can indicate


PLF by detecting abnormalities of the stapes footplate (Weissman et al. 1994) or air bubbles at the
prosthesis following stapedectomy (Woldag et al.
1995). Magnetic resonance contrast imaging in
experimental PLF (in the cat) documented significant increases in signal intensity of the cochlear perilymph with pooling of enhanced perilymph in the
ipsilateral mastoid bulla (Morris et al. 1993).
Electronystagmography

Electronystagmographic recordings of thermal irrigation may be helpful in so far as they detect peripheral vestibular dysfunction by a concomitant
unilateral hyporesponsiveness in about one-half of
the patients (Singleton et al. 1978; Thompson and
Kohut 1979; Love and Waguespack 1981). There is,
however, no pathognomonic or characteristic ENG
sign for fistula patients.
Hearing tests

Audiological testing cannot provide adefinite diagnosis of the presence of a fistula, even if positional
hearing tests (Fraser and Flood 1982) and the lowfrequency air-bone gap test are included. The usual
finding is of some non-specific sensorineural hearing loss, lying between 5 and 10 dB, with fluctuating
high-pitched tinnitus and aural pressure. Electrocochleography was undertaken intraoperatively by
placing an electrode on the oval window and suctioning the round window. This caused a decrease in
action potential amplitude in two fistula patients
who had no visible leakage of fluid (Aso and Gibson
1994).
Exploratory tympanotomy

Exploratory tympanotomy is necessary to confirm the diagnosis. The low morbidity of surgical
exploration of suspected ears and the high percentage of positive findings lead authors to encourage its

use (Healy et al. 1974; Singleton et al. 1978; Love and


Waguespack 1981). The procedure can be carried out
under local anaesthesia. If the fistula is not readily
apparent, it can be demonstrated by having the
patient perform a Valsalva manoeuvre, by pumping
the stapes, or by lowering the patient's head
(Singleton et al. 1978). It is not necessary to confirm
the presence of a hole; observation of an abnormal
accumulation or welling-up of fluid is sufficient. In
the majority of oval window fistulas, the perilymph
leak is in the anterior portion of the stapes footplate.
It should be mentioned here that there is still a lively
debate going on between those who do and those
who do not believe in fistulas (Singleton and Weider
1987), focusing in particular on the meaning of the
phrase "surgically confirmed fistula."
Endoscopic diagnosis of PLF by using a
(transtympanic) needle scope and a (transtubal)
superfine flexible endoscope has had questionable
success (Ogawa et al. 1994; Wall and Rauch 1995).
Other proposed tests

The detection ofbeta-2-transferrin (a specific protein in the cerebrospinal fluid and perilymph) in the
middle ear was proposed as a diagnostic test, but it
has some limitations (Bordure et al. 1994; Thalmann
et al. 1994; Weber et al. 1994). A stained or coloured
perilymph would be a valuable tool for diagnosis of
PLF; however, intravenous fluorescein did not detect
experimental PLF in cats and dogs (Bojrab and
Bhansali 1993; Poe et al. 1993).

Differential diagnosis
The differential diagnosis of PLF includes other
forms of traumatic vertigo (p. 343): peripheral BPPV
(p.251) and central positional vertigo (p.291),
Meniere's disease (p. 83), vestibular paroxysmia
(p. 117), phobic postural vertigo (p. 469), "postconcussion syndrome" (p. 347), and bilateral
vestibulopathy (p. 127). It is particularly important
to consider PLF in children presenting with episodic
vertigo and sensorineural hearing loss, as well as in
patients who complain about vertigo and/or hearing
loss following trauma to the ear or head, or barotrauma (p. 351).

Aetiology and pathomechanisms


According to the various aetiologies of PLF, age of
onset may be throughout life with a peak in childhood

Perilymph fistulas

(congenital and acquired fistulas, p. 377) and a mean


age of 35 years for PLF in head trauma (Grimm et al.
1989). However, because of diagnostic uncertainties,
the incidence and prevalence of PLF are largely
unknown, as are those of bilateral fistulas (Kohut et
al. 1995).
PLF may be more common than previously recognised (Fee 1968; Black et al. 1991). In the only
histopathological study of PLF in humans to date,
Kohut et al. (1988) identified fistulas on histological
examination of the temporal bones in all 34 patients
whose clinical records showed symptoms of sensory
hearing loss and disequilibrium. The fistulas were
found in the oval window (32%), the round window
(26%), and in both oval and round windows (42%)
(Kohut et al. 1988). In clinical studies, the oval window is the most common location of PLF (about
50%), either in the fissula ante fenestrum area
anterior to the oval window or at the margin of the
footplate where the annular ligament has been disrupted (Black et al. 1991).
Labyrinthine fistulas can be classified by their

Ioeation (the oval window is more susceptible to


rupture than the round window) and
2. aetiology (traumatic, surgical, inflammatory,
tumourous, congenital).
1.

The perilymphatic space communicates with the


middle ear via the labyrinthine windows, and with
cerebrospinal fluid via the cochlear aqueduct and
the internal auditory canal (Fig. 23.1). They constitute the three major pathways of implosive and
explosive forces from middle to inner ear and from
inner to middle ear, respectively. Implosive damage
is produced by pressure applied via the Eustachian
tube. Explosive damage is produced by an increase
in the press ure of the cerebrospinal fluid, which can
result from violent exercise, heavy lifting, or even
sneezing (Sakikawa et al. 1994). Increased intracranial pressure is transferred to the inner ear via the
cochlear aqueduct, a bony channel found to be
patent in 100% of fetal bones but closed with fibrous
tissue in 70% of the examined bones of the elderly
(Wlodyka 1978). This mechanism may be the cause
of most of the so-called idiopathic fistulas, and the
age-dependent patency of the aqueduct would be in
good agreement with several reports on perilymph
fistulas in ehiIdhood presenting with hearing loss or
vertigo (Brockman 1959; Knight 1977; Grundfast
and Bluestone 1978; Supance and Bluestone 1983;
Petroff et al. 1986).
However, some of the fistulas in childhood are
congenital defects of the stapes footplate (Crook
1967; Rice and Waggoner 1967; Weider and Musiek
1984; Weber et al. 1993), which can be detected by

103

computed tomography in about one-third of the


cases (Weissman et al. 1994). Congenital and
acquired fistulas should be considered in children
who have sudden unexplained or progressive sensorineural hearing loss (Petroff et al. 1986), particularly in the presence of an enlarged vestibular
aqueduct and abnormal round window (Belenky et
al. 1993) but also in otitic meningitis (Althaus 1981).
Mondini dysplasia can be genetic, and, because of
frequent stapes deformities, perilymph fistulas and
meningitis are recognised as typical complications
(Illum 1972; Schuknecht 1980).
Perilymph fistulas (oval and round window rupture) as well as cochlear membrane rupture (hearing
loss, tinnitus) can result from barotrauma caused by
flying or diving (Pullen 1992) (see Barotrauma,
p. 351), during which considerable pressure differences occur between the inner and the middle ear
(Goodhill1971; Molvaer and Natrud 1979; Gussen
1981; Nakashima et al. 1988).
Traumatic oval window fistula was first described
by Fee (1968), who presented two cases. Further
cases were reported subsequently by Healy et al.
(1974). PLF and endolymphatic hydrops may
develop following transverse temporal bone fracture
(Lyos et al. 1995). It is now widely recognised
that oval and round window fistulas occur in
patients after concussive and non-concussive (mild)
head trauma (Grewal et al. 1983; Lehrer et al. 1984;
Grimm et al. 1989; Glasscock et al. 1992; Legent and
Bordure 1994). Even on rare occasions they follow
acoustic trauma, a blow to the external auditory
canal from a fist, manipulation of the ear (cleaning
cerumen with cotton swabs or sticks; Kubo et al.
1993), or water jet irrigation (Wurtele 1981; Anon
and Miller 1985).
Persistent oval window membrane fistulas following stapedectomy are well-known complieations of
surgery (Lewis 1961; Harrison et al. 1967; House
1967; Hemenway 1968; Goodhill1967), in particular
in the days of the polyethylene strut (Singleton and
Weider 1987). Air bubbles at the end of the prosthesis
can be imaged as indirect signs of fistula with highresolution computed tomography (Woldag et al.
1995). When the technique was changed to include
prostheses, fistulas became infrequent. Instead, PLF
as a complication of tubing or cochlear implant
surgery has made its appearance (Kubo et al. 1993).
Mastoidectomy and other otological procedures such
as surgery for cholesteatoma (which preferentially
damages the horizontal semicircular canal) or
gumma also carry the risk of a fistula.
Controversy surrounds the definition of
"spontaneous" PLF (Gibson 1993). Meyerhoff (1993)
reviewed 212 patients who underwent surgical
exploration for suspected fistula. Fifty-eight percent

104

had an antecedent history of an extern al event (trauma, flying, diving), while almost 41 % recalled an
antecedent event of internalorigin (lifting, straining,
sneezing, nose blowing). Thus, if a spontaneous
event is defined as occurring or produced by its own
energy, less than 2% were considered to have had
spontaneous PLF (Meyerhoff 1993). One should,
however, distinguish between the aetiological cause
of PLF and the trigger of its manifestation.

Experimental perilymph/endolymph fistulas


and endolymphatic hydrops
The inner ear pathology of experimental PLF was
studied in the guinea-pig by injecting artificial perilymph into the subarachnoid space or by suctioning
4 111 of perilymph through one of the round window
membranes (Nomura et al. 1992a). Secondary
pathology of the otolith organs and the semicircular
canals mainly consisted of a collapse of the membranous labyrinth with rupture of Reissner's membrane (Nomura et al. 1992a). The macula utriculi was
then covered by the collapsed wall (Figs. 6.3, 6.4),

Fig. 6.3. Histopathological findings of experimental perilymph


fistula in a guinea-pig showing a collapse of membranous
labyrinth (top), which touches the top of the lateral crista (top,
right) and part of the utricular macula (top, left). The trabecular
mesh is absent (these findings are similar to vestibular atelectasis,
p. OOO). The control (bottarn) shows normal membranous
labyrinth with trabecular mesh preserved. (From Nomura et al.
1992a.)

Vertigo
which can cause inadequate stimulation of the sensory epithelium (Nomura et al. 1992a; Kukita and
Nomura 1994), resulting in dizziness and unsteadiness, especially with head motion and pressure
changes. This condition is termed "floating"
labyrinth (Nomura et al. 1992b), since the moderately
collapsed membranous labyrinth may drift with
cerebrospinal fluid and/or perilymph pressure
changes, thereby stimulating sensory hair cells of the
utricle or semicircular canals (Fig. 6.5). Caloric
"irregularities" are seen in experimentally induced
PLF similar to those in patients (Young et al. 1992).
The recovery of caloric excitability indicates the
healing of the fistula (Young and Nomura 1995).
Endolymphatic hydrops and associated Meniere's
disease (p. 87) may develop secondary to perilymphatic fistulisation. Cochlear hydrops was observed
in guinea pigs after obliteration of the ductus
reuniens (Kimura et al. 1980), and after experimental
perilymph fistulas (Nomura et al. 1987). Cochlear
hydrops was also found in human temporal bone in
association with collapsed saccule and ductus
reuniens (Kitamura et al. 1982). On the other hand,
endolymph fistulas (defects of the membranous
labyrinth) occur in Meniere's disease (Schuknecht
1993; Koskas et al. 1983). An acute endolymph fistula
causes the vertigo attack (sudden rupture of the
membrane in endolymphatic hydrops), whereas a
permanent fistula will lead to permanent recovery
because it shunts the endolymphatic hydrops.
Fistulisation of various parts of the membranous
labyrinth was, therefore, used to treat Meniere's disease, and cochlear endolymphatic shunt procedures
reduced the magnitude of experimental hydrops in
animals (Kimura 1984).

Fig.6.4. Enlarged detail from Fig. 6.3, showing the collapsed


membranous labyrinth covering the utricular macula. (From
Nomura et al. 1992a.)

105

Perilymph fistulas

trabecular mesh

Fig.6.5. Schematic representation of normal utricle (top, right) and semicircular canal (top, left) in experimental perilymph fistula of
the guinea-pig. A collapse of the membranous labyrinth occurs (bottom). A collapsed labyrinth may drift (to -and-fro arrows) with
cerebrospinal fluid and / or perilymph pressure changes resulting in inadequate stimulation of part of the utricular nerve ceillayer
(/arge arrows). The "floating" labyrinth can cause dizziness and unsteadiness. (From Nomura et al. 1992b.)

Management
Management of PLF is either
1. conservative or
2. surgical.
In the acute case, conservative treatment is universally recommended, since most fistulas he al spontaneously. Immediate surgical intervention has also
been proposed by some authors (Pullen et al. 1979)
in cases where there is no doubt of the diagnosis
because of a history of barotrauma. As initial treatment, however, surgical interventions should be
avoided, because there is a considerable risk of postoperative recurrence of PLF, ranging between 10 and
47% (Black et al. 1991; Gyo et al. 1994; Singleton et al.
1978; Seltzer and McCabe 1986), and of a secondary
labyrinthine hydrops (Potter and Conner 1983;
Grimm et al. 1989). Furthermore, prolonged conser-

vative treatment cannot be replaced by surgery, since


strict postoperative bedrest and medicinal treatment
are required post -surgery.

Conservative treatment
Conservative care consists of absolute bedrest, with
the head elevated, for 5-10 days. Prolonged bedrest
of 6 weeks may be more effective, since collagen
healing requires 6-12 weeks for fractures and severe
sprains, aperiod in which skin achieves only a 50%
recovery of tensile strength (Harris 1979; Grimm et
al. 1989). There is general agreement on a good prognosis for vestibular symptoms in PLF (Healy et al.
1976; Seltzer and McCabe 1986; Shelton and
Simmons 1988), and hearing recovery can be
ensured if conservative treatment is started early in
PLF patients with mild hearing loss (Kubo et al.
1993). Avoidance of straining, sneezing, coughing,
loud noise stimulation, or head-hanging positions,

Vertigo

106

and the use of stool softeners are important to


reduce further explosive and implosive forces that
can activate perilymph leakage (Singleton et al. 1978;
Anon and Miller 1985). If symptoms clear, the
patient is sent horne and advised to limit physical
activity and to do no lifting or straining for another
2 weeks.

Surgical treatment
If symptoms persist for more than 4 weeks or if
hearing loss worsens, exploratory tympanotomy is
indicated. If a fistula is found (e.g. of the oval window), repair of the leak with tissue is preferred
before considering stapedectomy (Palva 1983; Anon
and Miller 1985; Singleton and Weider 1987). As fat
does not work weIl, small pie ces of perichondrium,
temporalis fascia, or periosteum are used. It is difficult to put tissue into the tiny crevice of an oval window fistula; it is easier to fix a leakage in the round
window because of the soft membrane behind it.
Modern techniques show that the use of laser to prepare the graft bed, multiple postauricular areolar
tissue grafts (Seltzer and McCabe 1986; Althaus
1981), otologous fibrin glue (Moretz et al. 1986), and
strict adherence to postoperative instructions
designed to minimise sudden changes in intracranial pressure can significantly decrease the PLF
recurrence rate to less than 10% (Black et al. 1991).
But even in the latter ENT setting only 15% of
patients with clinically suspected PLF underwent an
operation, whereas the majority were managed conservatively. Statistics on the efficacy of tympanoscopy with fistular repair are difficult to
interpret, since prophylactic grafting is recommended
for both oval and round windows even if no fistula
or perilymph leak is identified on tympanotomy
(Black et al. 1987, 1991; Simmons 1982; Weider and
Johnson 1988; PareIl and Becker 1986). In a strict
sense, this means that one cannot be sure in this
group of patients if a fistula existed beforehand and
was repaired.
The results of these surgical interventions are not
encouraging, especially with respect to improvement
of the hearing defect which is only on the order of
25-50% (Healy et al. 1974; Althaus 1977; Love and
Waguespack 1981; Seltzer and McCabe 1986;
Simmons 1982). Prognosis is better for vestibular
symptoms (Althaus and House 1973; Shelton and
Simmons 1988); disequilibrium and vertigo resolved
in 80-90% of patients post-surgically (Black et al.
1991;1992). The uncertainty of a possible recurrence
of PLF or development of a secondary hydrops is so
great, however, that it is still a subject of debate, for
example, whether an aeroplane pilot should be

allowed to continue flying after surgical repair of a


fistula.
If labyrinthine fistula is caused by cholesteatoma,
relief from fistula symptoms is possible by interrupting the semicircular canals and obliterating the fistula with otologous material in aI-stage open-method
tympanoplasty without damaging cochlear function
(Kobayashi et al. 1995). In otherwise untreatable PLF
syndromes the application of gentamicin or laserlabyrinthectomy may be ultimately indicated to
selectively destroy vestibular sensory hair cells while
preserving hearing (Nomura 1994).

Tullio phenomenon
Sound-induced vestibular symptoms such as vertigo,
nystagmus, oscillopsia, and postural imbalance in
patients with perilymph fistulas are commonly
known as the Tullio phenomenon (Tullio 1929). The
occurrence of a distressing "Tullio symptomatology"
presupposes PLF pathology; however, only rare
patients with PLF suffer from the Tullio phenomenon. It seems, nevertheless, justified to give a
detailed and separate description of the Tullio
phenomenon in conjunction with PLF, since this
pathological condition has revealed new details
about human vestibular function in connection with
ocular motor and postural control. Oculographic,
posturographic, and EMG studies allow a unique
analysis of vestibulo-ocular and vestibulospinal
otolith reflexes in humans using sound stimulation.

Experimental history
Deetjen (1899) and Richard (1916) were the first to
report vestibular symptoms during acoustic stimulation. In 1929, Tullio investigated this phenomenon
experimentally in pigeons, rabbits, chickens and
ducks, in which he performed fistulisation of the
bony labyrinth. In these animals, loud sounds
induced eye and head movements corresponding to
the semicircular canal stimulation. Tullio wrongly
assumed that the excitability of the cristae
ampullares represented an increased physiological
sound response, which is functionally significant for
"orientation sound reflexes". Huizinga (1934, 1935),
Huizinga et al. (1951), Jellinek (1928), Dohlman
(1931) and van Eunen et al. (1943), while confirming
the "Tullio phenomenon" in different species by fistulisation, did not agree with TuIlio's interpretation
of the underlying mechanical mechanism. They
argued that the sound waves could only be transmitted and thereby produce an excitation of the cristae

107

Perilymph fistulas

ampullares as a consequence of the pathological fenestration of the labyrinth.


Clinical history

In humans, two mechanisms are generally accepted


as causes of a Tullio phenomenon: in the first, more
than one mobile window or a fistula opens into the
vestibular labyrinth (Cawthorne 1956; Kacker and
Hinchcliffe 1970); in the second, a pathological contiguity of the tympano-ossicular chain and the
membranous labyrinth exists, for example if the
stapes is in contact with the otolith because of
endolymphatic hydrops (Cody et al. 1967; Kacker
and Hinchcliffe 1970; Brandt et al. 1988).
Clinically, single cases have been described in
association with labyrinthitis (Tullio 1929); canal
fenestration in the treatment of otosclerosis (Menzio
1952); haematotympanon in laterobasal fracture of
the skull (Spitzer and Ritter 1979); barotrauma
(Grundfast and Bluestone 1978) in a professional
horn player (Brandt et al. 1988; Dieterich et al. 1989);
perilymph fistulas of the round and oval windows
(Goodhill, 1967; Fee 1968; Grundfast and Bluestone,
1978; Grewal et al. 1983; Ildiz and Dundar 1994);
Meniere's disease (Naito 1955; Lange 1966; Kacker
and Hinchcliffe 1970); vestibular neuritis (Lange
1966); congenital temporal bone abnormalities
(Kwee 1972); and idiopathic cases (Bronstein et al.
1995).
(/inical types ofTullio phenomena

An otolith type can be differentiated within the heterogeneous group of Tullio phenomena, wh ich must
be distinguished from a semicircular canal (nystagmus) type, e.g. due to window rupture. In the latter,
the pathological elasticity of the bony labyrinth
makes it possible for high-intensity sound to move
the peri-endolymph system of the canals rather than
to push the otoliths. Click-evoked vestibulocollic
reflexes were studied in a patient with a unilateral
Tullio phenomenon who showed an abnormally low
threshold and larger re action when elicited from the
symptomatic side (Colebatch et al. 1994; Bronstein et
al. 1995). This is compatible with a pathological
increase in the normal vestibular sensation to sound.
Most of the older case descriptions in the literature
suffer from imprecise descriptions or the failure to
record induced eye/head movements, so that it is
impossible retrospectively to classify them as an
otolith or a semicircular canal type.

Otolith Tullio phenomenon


On the basis of an otoneurological examination of a
typical patient as well as re-evaluation of cases
described in the literature, evidence was presented
that an otolith Tullio phenomenon due to hypermobile stapes footplate typically manifests with the
pattern of sound-induced paroxysms of ocular tilt
re action (OTR) (Brandt et al. 1988; Fries et al. 1988;
Dieterich et al. 1989).
The patients complain about distressing attacks of
vertical oblique and rotatory oscillopsia (apparent
tilt of the visual scene, Fig. 6.6) as well as postural
imbalance (fall toward the unaffected ear and backward) elicited by loud sounds, particularly when
applied to the affected ear with a maximum at a certain frequency (e.g. 500 Hz). Uttering vowels or
blowing the no se causes similar symptoms of varying severity.
Clinically, simultaneous paroxysms of eye-head
synkinesis (ocular tilt reaction, OTR, p. 179) can be
observed with the triad of skew deviation (ipsilateral
over contralateral hypertropia), ocular torsion, and
head tilt toward the undermost eye. Electronystagmographic recordings as weIl as special video
analysis (time resolution: 1000 images/s) revealed a
latency for the eye movements of 22 ms with an
initial rapid and phasic rotatory-upward deviation
(Dieterich et al. 1989); this was followed by a smaller
tonic effect as long as sound stimulation las ted (Fig.
6.7). This short latency agrees with the short-latency
compensatory eye movements (16.4 to 18.5 ms)
found with brief periods of free fall in the monkey
(Bush and Miles 1996). Skew deviation in the patient
is caused by a disconjugate larger deviation of the
ipsilateral eye. Repetitive sound stimulation leads to
habituation of the phasic component of eye movements. Rottach et al. (1996) reported a latency of
16 ms for sound stimuli-induced horizontal-torsional
nystagmus in a patient with Tullio phenomenon.
Oscillopsia and vertical eye movements with a
longer latency of 2.2 s were described in another
patient by Cohen et al. (1995). In our patient, a surprisingly short-latency vestibulospinal reflex was
recorded electromyographically (Fries et al. 1988),
with an EMG response after 47 ms in the tibialis
anterior muscle and after 52 ms in the gastrocnemius
muscle during upright stance (Fig. 6.6). A considerable postural perturbation was measured by means
of a posturography platform with the shortest latency of 80 ms and a direction-specific diagonal body
sway. Increasing intracranial pressure by Valsalva
manoeuvre may evoke slow tonic eye movements

Vertigo

108
eye movements
latency: 22ms

EMG latenc y:

~
~
postU"aI sway
latency: - 80ms

lOmm..........

Fig. 6.6. An otolith Tullio phenomenon (Jeft) is characterised


bya sound-induced ocular tilt reaction (skew deviation with ipsilateral over contralateral hypertropia, ocular torsion counterclockwise, head tilt with ipsilateral ear up (top) and increased
body sway predominantly from right-backward to left-forward
(bottom). Latencies of disconjugated eye movements were
22 ms for the left eye; latencies for the vestibulospinal reflex at
upright stance were 47 ms in the left tibialis anterior muscle and
52 ms in the left gastrocnemius muscle (centre) . Measurable postural sway had a minimal latency of about 80 ms (bottom).
(Brandt et al. 1988.)

ing stapedectomy eauses an ipsilateral transient OTR


(Halmagyi et al. 1979). This specific role of the utricle
in the generation of OTR has been supported by
findings in animal experiments in cats (Suzuki et al.
1969) and in guinea-pigs (Curthoys 1987) using electrieal stimulation of single utricular nerves or
loealised eleetrieal stimulation of spots on the utrieular maeula, respectively. Synaptic organisation of
utricular input provides a pattern of activation for
both spinal motor neurons (head tilt, body sway)
and eonjugate eyclodeviation with disconjugate vertieal divergenee (Gacek 1971; Reisine and Highstein
1979; Lang et al. 1979; Carpenter and Cowie 1985).
Otolith Tullio phenomena may not be as rare as
originally thought. The two most obviously detailed
ease deseriptions of a Tullio phenomenon by Deecke
et al. (1981) and by Vogel et al. (1986) include typical
features of OTR, although they described the syndrome in different terms in their articles. The
patient deseribed by Deecke et al. (1981) exhibited
head tilt to the left with diseonjugate oeular torsion
to the left, lasting throughout an utterance. The
patient of Vogel et al. (1986) exhibited mainly vertical eye movements, composed of an initial component followed by a slower re setting movement,
wh ich was often divided into two parts with different velocities. These authors diseussed a possible
otolithie mechanism without, however, providing
surgieal proof of the site of the fistula. Another
patient, described by Spitzer and Ritter (1979) in
retrospeet, suffered from an otolith Tullio phenomenon, whieh in his case was due to fraeture of the
medial wall of the tympanon and involved the stapes
footplate, eausing sound-indueed eontraversive
head- and body-tilt without nystagmus (Fig. 6.8).

and oseillopsia opposite in direetion to those of the


Tullio phenomenon (Fig. 6.7).

Vestibulospinal reflexes tested as part of the


Tullio phenomenon

Why does otolith Tullio phenomenon manifest with


paroxysmal ocular tilt reaction (OTR)?

The sound-induced paroxysmal OTR permits


otolithie reflex studies whieh hitherto were not possible in healthy subjects. The functional role of the
otoliths in counteracting rapid perturbations of the
body in order to maintain postural balance is still
under discussion. It has been argued that the otolith
responds too late to counterbalance perturbations
(Dietz and Berger 1982; Diener et al. 1983), because
the vestibular system initiates muscle response only
after 180 ms (Nashner 1976). The authors believed
that these were mediated by somatosensory reflex
mechanisms.
On the other hand, there is evidenee from experiments with sudden unexpected falls (Melvill Jones
and Watt 1971; Greenwood and Hopkins 1976) as
well as perturbation studies in labyrinthectomised

The hypo thesis that the Tullio phenomenon arises


from non-physiologieal meehanieal otolith stimulation is based on (1) the loeation of the otoliths
direct1y adjacent to the stapes footplate and (2) the
typieal response pattern of OTR. Surgical exploration of the middle ear of our patient revealed a subluxated stapes footplate with the hypertrophie
stapedius muscle causing pathologically large amplitude movements during the stapedius reflex. OTR is
an eye-head synkinesis initiated by stimulation
(Westheimer and Blair 1975) or lesion (Brandt and
Dieterich 1987) of the otoliths or graviceptive pathways (p. 180). Inadvertent utricular damage follow-

Perilymph fistulas

109
EYE MOVEMENTS

SOUND
490 Hz/95 dB

OSCILLOPSIA

VERTICAL EOG

~\ ( ~
........'
,
~
, :. "
",/

':, . .....

.......
:.

"' :',

VALSALVA
MANOEUVRE

' f+
.. :
<@>
I

I..

~'

-J

-"

Fig.6.7. Tullio phenomenon of sound-induced eye movements due to a hypermobile stapes footplate (left earl. The sound causes a
rapid and phasic eye movement oblique upward with incyclotropia and concomitant oscillopsia with counterclockwise tilt of the visual scene. A smaller tonic deviation of the eyes continues as long as the sound lasts (top).lncreasing intracranial pressure by Valsalva
manoeuvre causes smaller slow and tonic eye movements and oscillopsia opposite in direction to the Tullio phenomenon (bottom).
The opposite directions of eye movements may reflect push or pull stimulation of the otoliths. EOG '" electro-oculogram. (Dieterich et
al. 1989.)

.~,,
<S2>
,,

''

~
~
J
Fig.6.8. Schematic drawing of the push and pull mechanism of the otoliths caused by tilting of the stapes footplate, which is
induced either by the stapedius reflex (push) or the Valsalva man oeuvre (pulI) in an otolith Tullio phenomenon.ln the latter condition
increased intracranial pressure is transmitted toward the middle ear. Direction of induced eye movements are opposite, either up or
down, indicative for the mode of stimulation. U '" utricle, 5", saccule.

animals (Watt 1976; Lacour and Xerri 1980) that the


otoliths eontribute signifieantly to the vestibulospinal reflex with muscle aetivatin after 60-80 ms.
The strang EMG response in our patient, which was
seen, at the earliest, after 47 ms and eaused a
measurable body sway after 80 ms, strongly suggests
the important funetion of vestibulospinal reaetions
transdueed by the otoliths. Calculations taking into
aeeount the delay of the stapedius reflex by 5-10 ms
(Salomon and Starr 1963) suggest that the reflex are
frm the utrieularlsaeeular maeula to the effeetor

muscle is based on a three-neuron are (see vestibulospinal reflexes, p. 10) for the earliest eomponent. The
22 ms lateney of vertieal eye movements is
eonsistent with the well-known three-neuron are
vestibulo-oeular reflex.
Vestibulospinal effeets in our studies were modulated to a greater degree by the patient's body position and were abolished with eyes open (Fig. 6.9) or
in supine position. The latencies were different for
agonists and antagonists, despite regular eoaetivatin. Inereased amplitude and different patterns of

Vertigo

110

activation of leg, arm, and neck muscles cannot be


interpreted as arising solely from three-neuron
reflex ares, but obviously involve preprogrammed
motor patterns intended to maintain postural stability.
The neck proprioceptive input also plays a role in
the vestibulospinal reflex: turning the head about the
vertical z-axis to the left or right and maintaining it
in that position does not alter the position of utriculus and sacculus with respect to the gravitational

tib. ant. m. lell

0 .5mv

eyes closed

0 .4

0 .3

0.2
0.1

0.5mv

100

200

300

400

500ms

field. Nevertheless, there are marked and reproducible differences in the latencies of lower leg
muscle activation. Turning the head to the right, i.e.
toward induced head tilt, results consistently in an
activation of the tibialis anterior muscle at about 50
to 60 ms, whereas voluntary turning of the head to
the left, i.e. against induced head tilt, increases
latencies to 80-87 ms (Fig. 6.10). Not only the onset
of muscle activity but also its peak in the rectified
EMG response occurs about 20-30 ms later under
head-turned-Ieft conditions. It appears therefore
that the somatosensory input from neck muscles can
modify the timing of muscular activation in the
lower leg following otolithic stimulation.
Both somatosensory and static otolithic inputs
are alte red when the patient rests on hands and
knees (crawling position) . The pattern of activation
is slightly decreased in amplitude, yet the latencies
do not differ significantly from those in the upright
stance. Extension or retroflexion of the head does
not alter the pattern of activation in a patient with
typical otolith Tullio phenomenon. Functional inactivation of vestibulospinal reflexes is dependent on
assumed posture (Fries et al. 1993). The effect of
otolithic stimulation was studied under several conditions in wh ich the lower leg muscles were voluntarily contracted but not used to maintain upright

eyes ope n

0 .5mv
0 .4

0.5mv

~
C1>

0.3

0.2

tib . an!. m. lell

'0

01
Q)

1:.

500ms

0.1
I

300

400

s60ms

eyes closed
tib. ant. m. lett

~
c

500ms

Fig.6.9. Otolith Tullio phenomenon: suppression of vestibulospinal reflex activity in the ipsilateral tibialis anterior muscle
when the patient fixa ted a stationary visual scene during sound
stimulation b.ln a and c the muscular response with eyes closed
immediately before and afterward is shown. (Fries et al. 1993.)

<:

.g
C1>

E
'0

01

C1>
1:.

Fig. 6.10. Influence of neck proprioceptions on latency of


vestibulospinal reflex recorded in the ipsilateral tibialis anterior
muscle following otolith stimulation. Turning the head to the left
a increases latencies to 80- 87 ms; turning the head to the right b
results in activation of muscle at 50- 60 ms. (Fries et al. 1993.)

111

Perilymph fistulas

posture. It was common to all these conditions that,


in spite of the continuous voluntary discharge in the
EMG, no increase (or decrease) of activity time
locked to the sound stimulation could be detected.
When the patient balances on one foot (eyes closed),
the tibialis anterior muscle in the supporting leg is
activated after a short latency; there is a slightly
more pronounced response in the left leg (ipsilateral
to stimulation) than in the right one (Fig. 6.11).
Conversely, the EMG of the elevated leg remains
unmodulated, showing a sustained discharge due to
voluntary contraction of the muscles. When the
patient is sitting or supine, no specific response can
be recorded from the lower leg muscles. Figure 6.12
illustrates that excitatory muscle activation in neck
and upper limb muscles following otolithic stimulation is delayed compared to the lower leg antigravity
muscles, in spite of the considerable difference in the
distance of their respective motor neurons from the
vestibular nuclei (Fries et al. 1993).
Otolithic stimulation, therefore, does not release a
rigid reflex but triggers different patterns of antigravity muscle activation, depending on the current
posture. This flexibility is necessary to maintain balance in situations with combined voluntary-active
and involuntary-passive stimulation, for example,
walking on a rolling ship.

<'CI

tib . ant. m. right

0 .5mv

0 .4

0;

0 .3

,~~wJ\~I'N 1~

Cl

21

]!

0.2

0. 1
0

0.5mv

In principle, the management of the Tullio phenomenon is identical to that of PLF (p. 105). Experience
in management of these particular patients is based
only on single case descriptions, such as surgical
stapedectomy (Lange 1966), cutting of the stapedius
muscle or surgical fixation of the stapes footplate by
interposition of cartilage (which resulted only in a
transient relief in our patient), or compressed silastic foam inserted between anterior and posterior
crus of the stapes so that when it expands, it fixes the
stapes within the middle ear (Dieterich et al. 1989).
The percentage of spontaneous recoveries with
medical therapy is unknown, but surgical interventions

Management

tib. ant. m. lett

0 .5mv

'0

Quantitative measurements of postural sway


(posturography) with and without high-intensity
sound stimulation have been proposed as a functional test for diagnosis of PLF, because of its surprising sensitivity (Pyykk et al. 1992), but the
specificity of the test is still unsatisfactory. Patients
with Meniere's disease, chronic otitis media, or vertigo
of peripheral vestibular origin also exhibit increased
postural sway with sound stimulation (Ishizaki et al.
1991; Pyykk et al. 1992; Hadj-Djilani 1991).

100

200

300

400

500ms

tib . ant. m. lett

0 .5mv

500ms

tib. ant. m. right

'0
C

'"01

21

.<:

.~

500ms

Fig.6.11. Otolith Tullio phenomenon: suppression of vestibulospinal reflex activity in the tibialis anterior muscle of both sides when
not used for support of upright stance while the patient balanced on one foot. In the elevated leg the EMG remained unmodulated.
a Balancing on left foot, b balancing on right foot (Fries et al. 1993.)

Vertigo

112

should only be considered when the Tullio phenomenon is distressing and lasts from weeks to months.

neck m I t1

25mv

20
15

.05
0

Fistula of the anterior semicircular


canal

10

300

400

500ms

troceps m lelt

0 .5mv

OA
0 .3
0 .2
0. 1
0

SOOms

.2Smv

lorearm extensor left

O.Smv

t ib. an!. m. le ft

A new syndrome has been described by Minor et al.


(1998) in six patients: a dehiscence of the bone overlying the anterior semicircular canal. The clinical
manifestation of this syndrome was sound- and/or
pressure-induced dizziness with vertical and rotatory
oscillopsia or diplopia associated with verticaltorsional eye movements typical for excitation or
inhibition of the anterior semicircular canal of the
affected ear. The bony dehiscence can be identified
on temporal bone CT scan, a finding that we were
able to confirm (Fig. 6.13). Disabling disequilibrium
in two patients prompted Minor and co-workers to
plug the affected canal via the middle cranial fossa
approach. The symptoms improved in each case
(Minor et al. 1998).

500ms

Fig.6.12. Otolith Tullio phenomenon: simultaneously recorded


activity of ipsilateral neck a, triceps brachii b, forearm extensor c
and tibialis anterior muscles d following sound stimulation of the
patient's left ear. The shortest latency, i.e. that of the tibialis anterior muscles, is shown as a vertical dotted li ne for comparison.
(Fries et al. 1993.)

Fig.6.13. Fistula of the anterior semicircular canal. Coronal CT


scan of temporal bone in a patient with sound- and pressureinduced vertical-torsional eye movements and oscillopsia. Note
dehiscence of temporal bone overlying the anterior canal
(arrow).

Perilymph fistulas

References
Althaus SR (1977) Spontaneous and traumatic perilymph fistulas.
Laryngoscope 87:364-371
Allthaus SR (1981) Perilymph fistulas. Laryngoscope 91:538-562
Althaus SR, House HP (1973) Long-term results of perilymph fistula repair. Laryngoscope 83:1502-1509
Anon JB, Miller GW (1985) Perilymph fistula. South Med J
78:1454-1457
Aso S, Gibson WP (1994) Perilymphatic fistula with no visible leak
of fluid into the middle ear: a new method of intraoperative
diagnosis using electrocochleography.Am J OtoI15:96-100
Belenky WM, Madgy DN, Leider JS, Becker q, Hotaling AJ (1993)
The enlarged vestibular aqueduct syndrome (EVA syndrome).
Ear Nose Throat 72:746-751
Black FO, Lilly D], Nashner LM, Peterka RI, Pesznecker SC (1987)
Quantitative diagnostic test for perilymph fistulas. Otolaryngol
Head Neck Surg 96: 125-134
Black FO, Pesznecker S, Norton T, Fowler L, Lilly DJ, Shupert C,
Hemenway WG, Peterka RJ, Jacobson ES (1991) Surgical management of perilymph fistulas. A new technique. Otolaryngol
Head Neck Surg 117:641-648
Black FO, Pesznecker S, Norton T, Fowler L, Lilly DJ, Shupert C,
Hemenway WG, Peterka RJ, Jacobson ES (1992) Surgical management of perilymph fistulas: a Portland experience. Am J
Oto113:5
Bojrab DI, Bhansali SA (1993) Fluorescein use in the detection of
perilymphatic fistula: a study in cats. Arch Otolaryngol Head
Neck Surg 108:348-355
Bordure P, Delaroche 0, Beauvillian C, Legent F (1994) Perilymph
fistula: diagnosis by detection of perilymph in the middle ear
by beta-2-transferrin immunofixation. Ann Otolaryngol Chir
Cervicofac 111:180-184
Borries GV (1923) Vaskulre Labyrinthfistelsymptome. Mschr
Ohrenheilk 57:443
Brandt Th, Dieterich M (1987) Pathological eye-head coordination in roll: Tonic ocular tilt reaction in mesencephalic and
medullary lesions. Brain 110:649-666
Brandt Th, Dieterich M, Fries W (1988) Otolithic Tullio phenomenon typically presents as paroxysmal ocular tilt reaction. Adv
Oto-Rhino- Laryngol 42: 153-156
Brockman SJ (1959) An exploratory investigation of delayed progressive neural hypacusis in children. Arch Otolaryngol
70:340-356
Bronstein AM, Faldon M, Rothwell J, Gresty MA, Colebatch I,
Ludman H (1995) Clinical and electrophysiological findings in
theTullio phenomenon. Acta Otolaryngol (Stockh) Suppl
520:209-211
Bush GA, Miles FA (1996) Short-latency compensatory eye movements assoeiated with abrief period of free fall. Exp Brain Res
108:337-340
Carpenter MB, Cowie RJ (1985) Connections and oculomotor projections of the superior vestibular nucleus and cell group 'y'.
Brain Res (Amsterdam) 336:256-287
Cawthorne T (1956) Chronic adhesive otitis. J Laryngol Otol
70:559-564
Cody DTR, Simonton KM, Hallberg OE (1967) Automatic repetitive decompression of the saccule in endolymphatic hydrops
(Tack operation). Laryngoscope 77: 1480-150 1
Cohen H, Allen JR, Congdon SL, Jenkins HA (1995) Oscillopsia
and vertical eye movements in Tullio's phenomenon. Arch
Otolaryngol Head Neck Surg 12:459-462
Colebatch JG, Rothwell JC, Bronstein A, Ludmann H (1994) Clickevoked vestibular activation in the Tullio phenomenon. J
Neurol Neurosurg Psychiatry 57:1538-1540
Crook JP (1967) Congenital fistula in the stapedial footplate.
South Med J 60: 1168-1170

113
Curthoys PD (1987) Eye movements produced by utricular and
saccular stimulation. Aviat Environ Med 58 (SuppI9) A: 192-197
Daspit CP, Churchill D, Linthicum FH (1980) Diagnosis of perilymph fistula using ENG and impedance. Laryngoscope
90:217-223
Deecke L, Mergner T, Plester D (1981) Tullio phenomenon with
torsion of the eyes and subjective tilt of the visual surround.
Ann NY Acad Sei 374:650-655
Deetjen H (1899) Akustische Strungen der Perilymphe. Z Biol
14:159-166
Diener HC, Bootz F, Dichgans J, Bruzek W (1983) Variability of
postural reflexes in man. Exp Brain Res 52:423-428
Dieterich M, Brandt Th, Fries W (1989) Otolith function in man:
Results from a case of otolith Tullio phenomenon. Brain
112:1377-1392
Dietz V, Berger W (1982) Spinal co ordination of bilateral leg
muscle activity during balancing. Exp Brain Res 47: 172-176
Dohlman G (1931) Diskussionsbemerkung. Acta Otolaryngol
(Stockh) 15:322
Fee GA (1968) Traumatic perilymph fistulas. Arch Otolaryngol
88:477-480
Fraser JG, Flood LM (1982) An audiometric test for perilymph fistula. J Laryngol Otol 96:513-520
Fries W, Dieterich M, Brandt Th (1988) Otolithic control of posture: Vestibulo-spinal reflexes in a patient with a Tullio
phenomenon. Adv Oto- Rhino-Laryngol 41: 162-165
Fries W, Dieterich M, Brandt T (1993) Otolith contributions to
postural control in man: short latency motor responses following sound stimulation in a case of otolithic Tullio phenomenon.
Gait & Posture 1:145-153
Gacek RR (1971) Anatomical demonstration of the vestibuloocular projections in the cat. Laryngoscope 81:1559-1595
Gibson WP (1993) Spontaneous perilymphatic fistula: electrophysiologic findings in animals and man. Am J OtoI14:273-277
Glasscock ME, Hart MJ, Rosdeutscher JD, Bhansali SA (1992)
Traumatic perilymphatic fistula: how long can symptoms persist. A follow-up report. Am J Otol13:333-338
Goodhill V (1967) The conductive loss phenomena in poststapedectomy perilymph fistula. Laryngoscope 77: 1179-1190
Goodhill V (1971) Sudden deafness and round window rupture.
Laryngoscope 81:1462-1474
Goodhill V, Harris I, Brockman SI, Hantz 0 (1973) Sudden deafness and labyrinthine window ruptures, audio-vestibular
observations. Ann Otol Rhinol Laryngol 82:2-12
Greenwood R, Hopkins A (1976) Muscle responses during sudden
falls in man. J PhysioI254:507-518
Grewal DS, Hiranandani NL, Pusalkar AG (1983) Traumatic perilymph fistulae of the round and oval windows. J Laryngol Otol
97:1149-1155
Grimm RJ, HemenwayWG, Lebray PR, Black FO (1989) The perilymph fistula syndrome defined in mild head trauma. Almquist
&Wiksell Tryckeri, Uppsala.
Grundfast KM, Bluestone CD (1978) Sudden and fluctuating hearing loss and vertigo in children due to perilymph fistula. Ann
OtoI87:761-771
Gussen R (1981) Sudden hearing loss assoeiated with cochlear
membrane rupture. Arch Otolaryngol107:598-600
Gyo K, Kobayashi T, Yumoto E, Yanagihara N (1994) Postoperative
recurrence of perilymph fistulas. Acta Otolaryngol (Stockh)
Suppl 514:59-62
Hadj-Djilani AMT (1991) Ataxia induced by acoustic stimulation
on force platform: results on patients with hearing loss and/or
vestibular lesion. Acta Otolaryngol (Stockh) SuppI481:447-450
Halmagyi GM, Gresty MA, Gibson WPR (1979) Ocular tilt reaction with peripheral vestibular lesion. Ann NeuroI6:80-83
Harris D (1979) Healing of the surgical wound. J Am Acad
Dermatoll:197-217
Harrison WH, Shambaugh GE, Derlaki EL, Clemis JD (1967)
Perilymphatic fistulas in stapes surgery. Laryngoscope
77:836-849

114
Healy GB, Strong MS, Feldman RG (1973) Ataxia secondary to
labyrinthine fistula. Laryngoscope 83:502-507
Healy GB, Strong MS, Sampogna 0 (1974) Ataxia, vertigo, and
hearing loss. A result of rupture of inner ear window. Arch
OtolaryngoI100:130-135
Healy GB, Friedman JM, Strong MS (1976) Vestibular and auditory
findings of perilymph fistula: a review of 40 cases. Otolaryngol
Head Neck Surg 82: 44-49
Hemenway WF (1968) Post-stapedectomy perilymph fistulas in
Rocky Mountain areas. Laryngoscope 78:1687-1715
Henneber! C (1905) Reflexe oto-oculo-moteur. Int Zib Ohrenheilk
3:405
House HP (1967) The fistula problem in otosclerotic surgery.
Laryngoscope 77:1410-1426
Huizinga E (1934) ber die Schallreflexe von Tullio. Pflgers Arch
234:665
Huizinga E (1935) On the sound reaction of Tullio. Acta
Otolaryngol (Stockh) 22:359
Huizinga E, de Vries HL, Vrolijk JM (1951) Analysis of the microphonic activity of the labyrinth of the pigeon into the contributions of various parts. Acta Otolaryngol (Stockh) 39:372
Ildiz F, Dundar A (1994) A case ofTullio phenomenon in a subject
with oval window fistula due to barotrauma. Aviat Space
Environ Med 65:67-69
Illum P (1972) The Mondini type of cochlear malformation. Arch
Otolaryngol 96:305-311
Ishizaki H, Pyykk I, Aalto H, Starck J (1991) Tullio phenomenon
and postural stability: experimental study in normal subjects
and patients with vertigo. Ann Otol Rhinol Laryngol
100:976-983
Jellinek A (1928) Akustische Reflexe an Tauben nach isolierter
Verletzung der knchernen Bogengnge. Mschr Ohrenheilk
62:241
Kacker SK, Hinchcliffe R (1970) Unusual Tullio phenomena. J
Laryngol Otol 84:155-166
Kimura RS (1984) Fistulae in the membranous labyrinth. Ann
Otol Rhinol Laryngol 93:36-43
Kimura RS, Schuknecht HF, Ota CY, Jones 00 (1980) Obliteration
of the ductus reuniens. Acta Otolaryngol (Stockh) 89:295-309
Kitamura K, Schuknecht HF, Kimura RS (1982) Cochlear hydrops
in association with collapsed saccule and ductus reuniens. Ann
OtoI9:5-13
Knight NJ (1977) Severe sensorineural deafness in children due to
perforation of the round window membrane. Lancet ii: 1003
Kobayashi T, Sato T, Toshima M, Ishidoya M, Suetake M, Takasaka
T (1995) Treatment of labyrinthine fistula with interruption of
the semicircular canals. Arch Otolaryngol Head Neck Surg
121:469-475
Kohut RI, Hinojosa R, Ryu JH (1988) Perilymphatic fistulae: a
single-blind
clinical
histopathological
study.
Adv
Otorhinolaryngol42: 148-152
Kohut RI, Hinojosa R, Thompson JN, Ryu JH (1995) Idiopathic
perilymphatic fistulas. A temporal bone histopathologic study
with clinical, surgical and histopathologic correlations. Arch
Otolaryngol Head Neck Surg 121:412-420
Koskas HJ, Linthicum FH, House WF (1983) Membranous ruptures in Meniere's disease: Existence, location and incidence.
Otolaryngol Head Neck Surg 91:61-67
Kubo T, Kohno M, Naramura H, Itoh M (1993) Clinical characteristics and hearing recovery in perilymphatic fistulas of different
etiologies. Acta Otolaryngol (Stockh) 113:307-311
Kukita N, Nomura Y (1994) Morphological changes of the vestibular labyrinth by experimental perilymph fistula. Showa Univ J
Med Sei 6:97 -103
Kwee HL (1972) A case of Tullio phenomenon with congenital
middle-ear abnormalities. ORL (Basel) 34:145
Lacour R, Xerri C (1980) Compensation of postural reactions to
free-fall in the vestibular neurectomised monkey. Exp Brain Res
40:103-110

Vertigo
Lang W, Bttner-Ennever JA, Bttner U (1979) Vestibular projections to the monkey thalamus: an autoradiographic study. Brain
Res 177:3-17
Lange G (1966) Das Tullio-Phnomen und eine Mglichkeit seiner
Behandlung. Arch Klin Exp Ohr Nas Kehlk Heilk 187:643-649
Legent F, Bordure P (1994) Post-traumatic perilymphatic fistulas.
Bull Acad Natl Med 178:35-44
Lehrer JF, Rubin RC, Poole DR, Hubbard JH, Wille R, Jacobs GB
(1984) Perilymphatic fistula - a definitive and curable cause of
vertigo following head trauma. West J Med 141:57-60
Lewis ML (1961) Inner ear complications of stapes surgery.
Laryngoscope 71:377-384
Love JT, Waguespack RW (1981) Perilymphatic fistulas.
Laryngoscope 91:1118-1128
Lucae A (1881) ber optischen Schwindel bei Druckerhhung im
Ohr. Arch Ohr Nas Kehlk Heilk 17:237
Lyos AT, Marsh MA, Jenkins HA, Cocker NJ (1995) Progressive
hearing loss after transverse temporal bone fracture. Arch
Otolaryngol Head Neck Surg 121:795-799
Melvill Jones G, Watt DG (1971) Muscular control oflanding from
unexpected falls in man. J PhysioI219:729-737
Menzio P (1952) I riflessi di Tullio in sogetti operati di fenestrazione labirintica. Otol ecc ItaI20:168, cited in Zbl Hals Nas
Ohren Heilk 66:374 (1953)
MeyerhoffWL (1993) Spontaneous perilymphatic fistula: myth or
fact. Am J OtoI14:478-481
Minor LB, Solomon 0, Zinreich JS, Zee OS (1998) Sound- and/or
pressure-induced vertigo due to dehiscence of the superior
semicircular canal. Arch Otolaryngol Head Neck Surg
124:249-258
Molvaer OI, Natrud E (1979) Ear damage due to diving. Acta
Otolaryngol (Stockh) SuppI360:187-189
Moon CN, Hahn M (1978) Pneumatic otoscopy and impedance
studies in middle ear diagnosis. Laryngoscope 88:1439-1448
Moretz WH Jr, Shea JJ Jr, Emmett JR, Shea JJ III (1986) A simple
autologous fibrinogen glue for otologic surgery. Otolaryngol
Head Neck Surg 95:122-124
Morris MS, Kil J, Carvlin MJ (1993) Magnetic resonance imaging
of perilymphatic fistula. Laryngoscope 103:729-733
Mygind SH (1918) Ein neues Fistelsymptom. Mschr Ohrenheilk
54:260
Naito T (1955) Three cases of Meniere's disease showing Tullio's
reaction. Otol Fukuoka 1:249 cited in Zbl Hals Nas Ohren Heilk
54:265 (1955/56)

Nakashima T, Itoh M, Sato M, Sato M, Watanabe Y, Yanagita N


(1988) Auditory and vestibular dis orders due to barotrauma.
Ann Otol Rhinol LaryngoI97:146-152
Nashner LM (1976) Adapting reflexes controlling the human posture. Exp Brain Res 26:59-72
Nomura Y (1994) Perilymph fistula: cancept, diagnosis and management.Acta Otolaryngol (Stockh) SuppI514:52-54
Nomura Y, Hara M, Funai H, Okuno T (1987) Endolymphatic
hydrops in perilymphatic fistula. Acta Otolaryngol (Stockh)
103:469-476
Nomura Y, Hara M, Young YH, Okuno T (1992a) Inner ear morphology of experimental perilymphatic fistula. Am J Otol
13:32-37
Nomura Y, Okuno T, Hara M, Young YH (1992b) "Floating"
labyrinth: pathophysiology and treatment of perilymph fistulas. Acta Otolaryngol (Stockh) 112:186-191
Nylen CO (1923) The labyrinthine fistula symptoms and pseudofistula symptoms in otitis. Acta Otolaryngol (Stockh) Suppl. 3
Ogawa K, Kanzaki J, Ogawa S, Tsuchihashi N, Inoue Y, Yamamoto
M (1994) Endoscapic diagnosis of idiopathic perilymph fistula.
Acta Otolaryngol (Stockh) SuppI514:63-65
Ostrowski VB, Hain TC, Wiet RJ (1997) Pressure-induced ocular
torsion. Arch Otolaryngol Head Neck Surg 123:646-649
Palva T (1983) Treatment of ear with labyrinth fistula.
Laryngoscape 93:1617-1619

Perilymph fistulas
PareIl GI, Becker GD (1986) Results of surgical repair of inapparent perilymph fistulas. Otolaryngol Head Neck Surg 95:344-346
Petroff MA, Simmons FB, Winzelberg 1(1986) Two emerging periIymph fistula "syndromes" in children. Laryngoscope
96:498-501
Poe DS, Gadre AK, Rebeiz EE, Pankratov MM (1993) Intravenous
fluorescein for detection of perilymphatic fistulas. Am I Otol
14:51-55
Potter CR, Conner GH (1983) Hydrops following perilymph fistula
repair. Laryngoscope 93:810-812
Pullen FW (1992) Perilymphatic fistula induced by barotrauma.
Am I Otol13:270-272
Pullen FW, Rosenberg GT, Cabeza CH (1979) Sudden hearing loss
in divers and fliers. Laryngoscope 86:1373-1377
Pyykk 1, lshizaki H, Aalto H, Starck I (1992) Relevance of the
Tullio phenomenon in assessing perilymphatic leak in vertiginous patients. Am I Otol13:339-342
Reisine H, Highstein SM (1979) The ascending tract of Deiters'
conveys a head velocity signal to medial rectus motoneurons.
Brain Res 170:172-176
Rice WI, Waggoner LG (1967) Congenital cerebro-spinal fluid
otorrhea via a defect in the stapes footplate. Laryngoscope
77:341-349
Richard D (1916) Untersuchung ber die Frage, ob Schallreize
adquate Reize fr den Vorhofbogengangsapparat sind. Z Biol
66:479-505
Rottach KG, Maydell RD von, DiScenna AO, Zivotofsky AZ,
Averbuch-Heller L, Leigh RI (1996) Quantitative measurements
of eye movements in a patient with Tullio phenomenon. I Vestib
Res 6:255-259
Salomon G, Starr A (1963) Electromyography of middle ear
muscles in man during motor activities. Acta Neurol Scand
39:161
Sakikawa Y, Kobayashi H, Nomura Y (1994) Changes in cerebrospinal fluid pressure in daily life. Ann Otol Rhinol Laryngol
103:959-963
Schuknecht HF (1993) Pathology of the ear. 2nd Ed, Harvard
University Press, Cambridge, Massachusetts
Schuknecht HF (1980) Mondini dysplasia. A clinical and pathological study. Ann Otol Rhinol Laryngol Suppl65, 89: 1-23
Seltzer S, McCabe BF (1986) Perilymph fistula: the lowa experience. Laryngoscope 94:37-49
Shelton C, Simmons FB (1988) Perilymph fistula: the Stanford
experience. Ann Otol Rhinol LaryngoI97:105-108
Simmons FB (1982) Perilymph fistula: Some diagnostic problems.
Adv Oto-Rhino-LaryngoI28:67-72
Singleton G, Weider D (1987) Panel discussion: Perilymphatic fistula. Am I OtoI8:355-363
Single ton GT, Karlan MS, Post KN, Bock DG (1978) Perilymph fistulas. Diagnostic criteria and therapy. Ann Otol Rhinol
LaryngoI87:1-7
Spitzer H, Ritter K (1979) Ein Beitrag zum Tullio-Phnomen.
Laryng Rhinol 58:934-936
Stenger HH (1953) Puls synchroner Pendelnystagmus.

115
Fistelsymptome ohne Fistel und Lagefistelsymptom. Arch Klin
Exp Ohr Nas Kehlk Heilk 162:213-228
Supance IS, Bluestone CD (1983) Perilymph fistulas in infants and
children. Otol Head Neck Surg 91:663-671
Suzuki 11, Tokumasu K, Goto K (1969) Eye movements from single
utricular nerve stimulation in the cat. Acta Otolaryngol
68:350-36
Thalmann 1, Kohut Rl, Ryu I, Comegys TH, Senarita M, Thalmann
R (1994) Protein profile of human perilymph: in search of
markers for the diagnosis of perilymph fistula and other inner
ear disease. Otolaryngol Head Neck Surg 111:273-280
Thompson IN, Kohut Rl (1979) Perilymph fistulae: Variability of
symptoms and results of surgery. Otolaryngol Head Neck Surg
87:898-903
Tullio P (1929) Das Ohr und die Entstehung der Sprache und
Schrift. Urban & Schwarzenberg, Munich.
van Eunen AIH, Huizinga HC, Huizinga E (1943) Die Tulliosche
Reaktion im Zusammenhang mit der Funktion des Mittelohrs.
Acta Otolaryngol (Stockh) 31:265
Vogel P, Tackmann W, Schmidt FI (1986) Observations on the
Tullio phenomenon. I NeuroI233:136-139
Wall C, Rauch SD (1995) Perilymph fistula pathophysiology.
Otolaryngol Head Neck Surg 112:145-153
Watt DGD (1976) Responses of cats to sudden falls: An otolith
originating reflex assisting landing. I NeurophysioI39:257-265
Weber PC, Perez BA, Bluestone CD (1993) Congenital perilymphatic fistula and associated middle ear abnormalities.
Laryngoscope 103:160-164
Weber PC, Kelly RH, Bluestone CD, Bassiouny M (1994) Beta
2-transferrin confirms perilymphatic fistula in children.
Otolaryngol Head Neck Surg 110:381-386
Weider D), lohnson GD (1988) Perilymphatic fistula: a New
Hampshire experience. Am I OtoI9:184-196
Weider D), Musiek FE (1984) Bilateral congenital oval window
microfistulae in a mother and son. Laryngoscope 94: 1455-1458
Weissman IL, Weber PC, Bluestone CD (1994) Congenital perilymphatic fistula: computed tomography appearance of middle ear
and inner ear abnormalities. Otolaryngol Head Neck Surg
111 :243-249
Westheimer G, Blair SM (1975) The ocular tilt reaction - a brainstem oculomotor routine. luvest OphthalmoI14:833-839
Wlodyka I (1978) Studies on cochlear aqueduct patency. Ann Otol
87:22-28
Woldag K, Meister EF, Kosling S (1995) Diagnosis in persistent
vertigo after stapes surgery. Laryngorhinootologie 74:403-407
Wurtele P (1981) Traumatic rupture of the eardrum with round
window fistula. I Otolaryngoll0:309-312
Young YH, Nomura Y, Hara M (1992) Caloric irregularity in experimentally induced perilymphatic fistula. Eur Arch
Otorhinolaryngol 249: 181-184
Young YH, Nomura Y (1995) Recovery of caloric function in
experimental perilymph fistula. Ann Otol Rhinol Laryngol
104:484-487

Peripheral vestibular paroxysmia


(disabling positional vertigo)

Episodic vertigo and other vestibular syndromes can


result from pathological excitation of various
vestibular structures: the labyrinthine receptors, the
vestibular nerve, the vestibular nudei, and their
ascending pathways to the thalamus and the cortex
(see Paroxysmal vertigo, p. 29; Table 2.1). There is
evidence that neurovascular cross-compression of
the eighth nerve is the probable cause of vestibular
paroxysmia (also termed disabling positional vertigo),
induding both paroxysmal hyperactivity and progressive functionalloss. Analogously to trigeminal
neuralgia, vestibular paroxysmia is diagnosed by the
occurrence of short attacks and series of rotation al
or to-and-fro vertigo, which are precipitated or
modulated by changing head position and frequently
associated with hypacusis and tinnitus (Brandt and
Dieterich 1994a).
This variable syndrome was first described by
Jannetta (1975) and later labelIed "disabling positional vertigo" by the same authors (Jannetta et al.
1984; M011er et al. 1986). The term covers a very
heterogeneous collection of signs and symptoms,
whieh can even mimick those of Meniere's disease
(M011er 1988). As important as this first description
of the condition was for the differential diagnosis of
episodic vertigo, as adefinition of a reliably diagnosable disease entity it nevertheless remains unsatisfying. Most reports on so-called disabling positional
vertigo suffer from the fallacious assumption that <Ca
certain kind of balance dis order that does not fulfil
the criteria of the established syndromes just discussed (Meniere's disease, benign paroxysmal positional vertigo, vestibular neuritis)" can be diagnosed
as a neurovascular cross-compression syndrome.
There are, of course, many more vertigo syndromes
relevant for differential diagnosis, and the short ca se
histories in these publications suggest that several
patients possibly suffered from vertigo diseases
other than neurovascular compression, such as
otolith dysfunction or secondary (somatoform)
phobie postural vertigo (see Chap. 32, p.469).
Jannetta et al. (1984) reported that symptoms in

three of nine patients were initiated by head trauma


or a benign paroxysmal positional vertigo. Phobie
postural vertigo (Brandt et al. 1994), panic attacks
with agoraphobia (Jacob et al. 1985) and space
phobia (Marks 1981) have been described as sequelae
of a vestibular disorder. A vestibular disorder initiated
the condition in about 20% of our patients with
phobie postural vertigo.
Lacking a well-defined syndrome and a diagnostic test, the non-surgical dinician finds it difficult to
believe in this disease. The increasing number oE
reports on "disabling positional vertigo" prompt us
to share our own preliminary experience with
episodic vertigo, a treatable condition that has long
escaped notice. The conservative therapeutic
approach to the typieal clinical syndrome of
neurovascular compression oE the eighth nerve is
mainly based on the eEficacy oE treatment with carbamazepine, the recurrence of attacks following
drug washout phases and the exdusion of a central
(particularlya demyelinating) disease (p. 242). There
is still no pathognomonic sign of the condition, and
to date the current imaging techniques for identifying causative nerve-vessel contacts leave much to be
desired (Fig. 7.1), since vessel contacts can also be
imaged (even at root entry zones) in asymptomatic
patients.

The clinical syndrome


Neurovascular cross-compression of the vestibular
nerve was the most probable diagnosis in our
patients, who had brief and frequent attacks of vertigo
(Brandt and Dieterich 1994a,b). According to our
findings, the diagnosis can be established on the
basis of six characteristic features:
1. short attacks of rotational or to-and-fro vertigo

117

lasting for seconds to minutes,

118

Vertigo
Table 7.1.

Peripheral vestibular paroxysmia (disabling position al verti-

go)
Clinical syndrome
Combination of short and frequent vertigo attacks and progressive
functionalloss of eighth nerve
- Short attacks of rotational or to-and-fro vertigo
- Attacks frequently triggered or modified by particular head positions
- Auditory or vestibular deficits and/or tinnitus
- Efficacy of carbamazapine
Incidence/age/sex
Rare condition that manifests throughout life (mean age 45 years)
without preference of sex
Pathomechanism
Analogously to trigeminal neuralgia, neurovascular cross-compression
of the root entry zone of the eighth nerve with local demyelination,
axonal hyperactivity and transversaily spreading ephaptic activation

Aetiology
Nerve compression by anterior or posterior inferior cerebeilar artery,
ectatic vein, vertebrobasilar dolichoectasia, tumour or bone

b
Fig.7.1. Neurovascular cross-compression ofthe eighth nerve.
a Axial projection, T2 -weighted MRI3D-CI55.The anterior inferior
cerebellar artery (AICA) can be delineated in the cerebelloponti ne angle in dose contact with the vestibulocochlear nerve
(arrow). b Sagittal projection of the internal acoustic canal, T2 weighted MRI 3D-CI55. The AICA crosses under the vestibular
nerve and there is nerve-vessel contact (arrow). (From Jger et al.
1997.)

Course / prognosis
Lasting for years to decades with varying frequency of attacks and
slowly progressive auditory and vestibular functionalloss
Management
Drugs: antiepileptic drugs (carbamazepine or phenytoin) or pimozide
Surgery: microvascular decompression ofthe eighth nerve
Differential diagnosis
- Basilar migraine
- Vertebrobasilar ischaemic attacks
- Atypical Meniere's disease
- Central vestibular paroxysmia (e.g. multiple sclerosis)
- Vestibular epilepsy
- Perilymph fistulas
- Benign paroxysmal positioning vertigo
- Phobic postural vertigo
- Bilateral vestibulopathy

2. attacks frequently provoked by particular head


positions and whose duration is modified by
changing head position,
3. hypacusis or tinnitus permanently or during the
attack,
4. auditory or vestibular deficits measurable by
neurophysiological methods,
5. efficacy of carbamazepine, and
6. a central cause excluded by clinical, neurophysiological and imaging investigations.

(ase reports

As distinct from typical trigeminal neuralgia - in


which there is no significant sensory loss - vestibular paroxysmia is characterised by paroxysmal
hyperactivity combined with some functional deficit
in the attack-free interval. Responsiveness to carbamazepine seems to be essential, since earlier studies
on "disabling positional vertigo" emphasised the
ineffectiveness of vestibular suppressant medications (M011er 1991a) without providing evidence
that antiepileptic drugs (first choice for trigeminal
neuralgia) had been systematically employed in
these patients.
Two case reports and a cumulative discussion of
typical signs and symptoms in suspected cases of
vestibular paroxysmia will attest to the importance
of the above key features (Brandt and Dieterich
1994b).

Case 1. The neurophysiological documentation on a


53-year-old man illustrates the subtle but measurable ocular motor and postural effects of the attacks
of vestibular paroxysmia (Fig. 7.2). The patient complained of oscillopsia, postural imbalance and unilateral tinnitus in the right ear during the 26 to
72-second-long attacks. These symptoms were provoked by head position. The upward spontaneous
nystagmus and ocular torsion of both eyes, as well as
pathological tilts of the subjective visual vertical
combined with auditory symptoms (unilateral highfrequency hearing loss, tinnitus at 2000 Hz) suggested
involvement of auditory, vertical semicircular canal
and otolith input. All together this pattern of perceptual, ocular motor and postural effects indicates a
paroxysmia of "graviceptive pathways" that subserve
vestibular function in the roll plane (see Chap. 10,

Peripheral vestibular paroxysmia (disabling positional vertigo)

L-f:.
---t: *
--t: -f:
---t: -f: ,,e
-t ---t:
eyes open

Inormal

Ipatient

(eyes closed)

head upright

eyes closed

NORMALS

l,o ~ 1

no attack

# 1

119

attack

RE

LE

VESTIBULAR PAROXYSMIA
no attack

head turn
to the right

during attack

head turn
to the lell

head
extension

0' - 3'

Fig.7.2. (a) Histograms for fore-aft (A-P) and lateral (R-L) postural sway obtained with a force-measuring platform (Kistler) in anormal
subject (top). The patient (No.l). who was suspected of having neurovascular cross-compression of the right eighth nerve, exhibited a
slightly increased body sway with eyes closed (bottom feft). This increased during the attack (bottom right), usually in a diagonal foreaft direction. Preferred body sway changed its direction by 90 if the head was turned to the right or to the left. (b) Schematic representation of the fundus, indicating normal position of the eyes in the roll plane (straight line through papilla-macula). Normal position
in roll is an excyclotropia of 0 to 10. In the patient (#1) eye position was slightly abnormal in both eyes (as seen by the observer).
During the attack a slightly clockwise rotation of 2_3 was repeatedly measured with laser-scanning ophthalmoscope. (From Brandt
and Dieterich 1994b.) RE = right eye; LE = left eye.

p. 175; Dieterieh and Brandt 1993; Brandt and


Dieterieh 1994b).
Case 2. Sinee ehildhood a 28-year-old man had
had frequent attaeks of foeal eontraetion of the left
frontalis muscle and oseillopsia due to a paroxysmal
skew deviation and torsional nystagmus (Straube et
al. 1994). During an attaek (20-80 s, 10-50 times per
day) the right eye moved downward, the left eye was
slightly elevated, both eyes were eyclorotated, and a
signifieant inerease of the eontraetion of the left
frontalis muscle elevated the eyebrow (Fig. 7.3). A
daily dose of 800 mg earbamazepine redueed the frequeney of the attaeks by more than 50%. The most
likely explanation for these signs is a peripheral vaseular eompression syndrome with paroxysmal diseharges of neurons of the peripheral vestibulooeular system and of faeial motor pathways. As in

ease 1, skew torsion of the eyes indieated excitation


of graviceptive pathways, sinee they were involved in
vestibular stabilisation of eye-head position in the
roll plane (Brandt and Dieterieh 1994e). A eombined
stimulation of the anterior semicireular eanal (AC)
and posterior semicireular eanal (PC) nerves of the
left side would eause a downward movement of the
right eye and an upward movement of the left eye,
combined with a cloekwise-beating torsional nystagmus (Cohen 1974). Isolated stimulation of the utrieular part of the vestibular nerve would eause a
similar skew deviation and torsion but no torsional
nystagmus (Suzuki et al. 1969). The eye movements
in this patient did not per mit us to differentiate
between aetivation of the AC and PC nerves or the
PC and utrieular nerve. However, these effeets are
best explained by a eombined stimulation of utrieular

Vertigo

120

or minutes. Nine of the 11 patients had single attacks


lasting 1 second to 4 minutes and two patients, 15
minutes to 2 hours (only one patient reported a continuous to-and-fro vertigo with postural imbalance
persisting for 2 to 3 days). As regards frequency of
the attacks, seven patients reported 1-30 per day,
two patients, 1-2 per week and the remaining two
patients had 1-3 per month. The frequency of the
attacks was relatively stable over months to years.
A change in head position was identified as the
provoking factor in eight patients. Changes included
extreme rotations (without relative change to the
gravitation al vector) and lateral head tiIts (with relative change to the gravitational vector). The particular manoeuvre of the head which provoked vertigo
was typical for each individual. However, its consistency in eliciting attacks was less than in benign
paroxysmal positioning vertigo (Chap. 16; p. 252).
Both intensity and duration of the attacks were successfully modified in five of the eight patients by
changing head position.
The attacks las ted long enough in five patients for
us to observe and register spontaneous and positional nystagmus, either vertical or horizontal in
direction. A vigorous rotational positional nystagmus with rotatory vertigo in one patient lasted undiminished for several minutes and led to vomiting.

Auditory symptoms and tests

(upward movement) and PC fibres (slightly downward movement) within the left vestibular and left
facial nerves (Straube et al. 1994). In fact, the fibres
of the utricular and facial nerves run adjacent to one
another in the proximal segment of the internal
auditory canal (Sando et al. 1972).
An analysis of the data for 11 patients diagnosed
as having vestibular paroxysmia revealed the following clinical spectrum (Brandt and Dieterich 1994b):

On the whole, the audiovestibular testing in 10 of 11


patients exhibited unilateral dysfunction in the symptom-free interval. Seven patients had either permanent and moderate impairment of hearing (n = 5)
and/or permanent tinnitus (n = 4), whereas only
three reported attacks of tinnitus accompanying the
vertigo. Pure-tone and speech audiometry were carried out. In accordance with the slowly progressive
hearing loss, slight to medium persistent abnormalities were found in five patients during the symptomfree inter val (high-frequency loss or deficits in the
middle-frequency range, 10-15 dB difference in
thresholds between both ears). The auditory nerve
conduction times were normal in the seven patients
in whom brain stern auditory-evoked potentials were
tested. Unilateral increased latencies between peaks I
and III have also been reported in patients suspected
to have disabling positional vertigo (M011er et al.
1986).

Vertigo

Other associated symptoms

Eight patients described their attacks as an apparent


rotation, two as to-and-fro vertigo and one reported
that to-and-fro vertigo lasted for hours with a superimposed rotational vertigo lasting for a few seconds

Other symptoms included the sensation of fullness in


one ear (n = 3), head tension (n = 3), postural imbalance (n = 3), oscillopsia (n = 2), and skew deviation
with ocular torsion. A single patient reported hearing

Fig. 7.3. Recurrent attacks with skew torsion and contraction of


the frontalis muscle in a 28-year-old man. Eye position between
the attacks a and during the attack b. The right eye moved downward and the left eye was slightly elevated. 80th eyes were
cyclorotated counterclockwise (slow phase).There was a significantly increased contraction of the left frontalis muscle during
the attack which induced a pronounced elevation of the eyebrow. (From Straube et al. 1994.)

Peripheral vestibular paroxysmia (disabling positional vertigo)

121

a cracking sound as the prelude to attacks of vertigo.


Apart from the above-described 11 patients, we have
also seen rare combinations of unilateral trigeminal
neuralgia, hemifacial spasm, positional vestibular
paroxysmia in extensive vertebrobasilar ectasia (Fig.
7.4), and vestibular paroxysmia with contraction of
the left frontal muscle in another patient (case 2; Fig.
7.3).

Electronystagmography

Whereas spontaneous and positional nystagmus


were observed only du ring the attacks, five patients
showed unilateral hyporesponsiveness to caloric
irrigation and one had a unilateral gaze-evoked
nystagmus in the symptom-free interval.
Subjective visual vertical

Perceived visual vertical (Dieterich and Brandt 1993)


was abnormal in four patients who showed tilted
adjustments from 2.9 to 6.0 (toward the ear
most probably affected) during the symptom-free
interval.
0

Posturography

The instability induced during the attacks can be


recorded as a direction-specific increased body sway,
particularly when the eyes are closed (Fig. 7.2).
The strongest argument for a peripheral nerve
origin of the condition of vestibular paroxysmia in
the discussed patients is based on the unilateral,
peripheral audiovestibular deficits documented in
most patients during a symptom-free interval. The
long duration of monosymptomatic attacks over a
mean of 7 years supports this argument, but it simultaneously makes an active (progressive) central
brainstem disorder less likely. The combination of
paroxysmal vestibular and auditory symptoms during an attack and their dependence on head position
are also indicative of a peripheral disorder. The
probable side with the affected nerve is indicated by
either a unilateral deficit of vestibular or cochlear
function in the symptom-free interval or by the
direction of eye movements induced during the
vestibular paroxysms (see Figs. 7.2 and 7.3).

Differential diagnosis
We have tried to delineate a largely homogeneous
group of patients who suffered from brief attacks of
vertigo and responded promptly and significantly to
carbamazepine. Basilar migraine can be excluded on
the basis of its clinical syndrome and the known

Fig. 7.4. Angiography of the left vertebral artery in a 45-yearold woman shows extensive ectasia of the vertebral and basilar
arteries. This vascular abnormality caused a combination of unilateral trigeminal neuralgia, hemifacial spasm and a positional
vestibular paroxysmia (mimicking benign paroxysmal positioning vertigo) most probably due to neurovascular compression of
three cranial nerves. (From Brandt 1991.)

ineffectiveness of carbamazepine. Some other possibilities, however, must be considered in a differential


diagnosis: atypical Meniere's disease, benign paroxysmal positioning vertigo, perilymph fistulas, bilateral vestibulopathy, vestibular epilepsy, non-epileptic
paroxysmal brainstem attacks (as described for multiple sclerosis (Andermann et al. 1959; Osterman
and Westerberg 1975, repetitive central vestibular
paroxysmia with vertigo and nystagmus (Lawden et
a1. 1995) and phobic postural vertigo (Chap. 32,
p. 469). None of our patients presented with the
characteristic features of phobic postural vertigo,
nor did they show an abnormal electroencephalogram (with focal slow activity or sharp waves over
temporoparietal regions), even when an occasional
attack occurred du ring the recording. Multiple
sclerosis had been specifically ruled out in advance
by careful diagnostic procedures (including MRI
scan, cerebrospinal fluid evaluation, visual and
somatosensory-evoked potentials). The 11 patients
described have been observed over the past 6 years.
During that time we have also seen another 14
patients, in whom the diagnosis of paroxysmal
brainstem attacks in multiple sclerosis (p. 244) was
convincingly established on the basis of associated

Vertigo

122

brainstem signs (dysarthria, ocular motor abnormalities, ataxia, central paresis) and typical findings
in neuroimaging and cerebrospinal fluid examinations. Hyperventilation may be clinically helpful for
provoking attacks in peripheral or central vestibular
paroxysmia, but this has not yet been studied systematically.

Aetiology and pathomechanisms


The age of our patients diagnosed as having vestibular paroxysmia ranged from 25 to 67 years (mean
age 51; mean duration of the condition before diagnosis 7.3 years; mean age of manifestation 44.5
years). There was no se"xual predominance (Brandt
and Dieterich 1994a).
Neurovascular cross-compression of the root
entry zone of the fifth, seventh and ninth cranial
nerves can cause local demyelinisation and axonal
hyperactivity (e.g. by transversally spreading ephaptic activation between bare axons and/or central
hyperactivity initiated and maintained by the
peripheral compression (Jannetta 1975; M0ller
1991 b) with the accompanying distressing symptoms of trigeminal neuralgia, hemifacial spasm and
glossopharyngeal neuralgia. Compressing vessels,
which had made indentations on the relevant nerve,
have been found surgically, and the efficacy of
microvascular decompression as a successful treatment of these symptoms is well established (Jannetta
1982; M0ller et al. 1986; M0ller 1991a). Ectopic
paroxysmal neuronal discharges with interaxonal
ephaptic transmission has been demonstrated for
hemifacial spasm (Nielsen 1984; Sanders 1989), in
(human) peripheral nerves (Tomasulo 1982) and in
animal experiments (Rasminsky 1980; Seltzer and
Devor 1979). It probably also accounts for the rare
condition of abducens neuromyotonia (HeImchen et
al. 1992; Barroso and Hoyt 1993).
These findings suggest that it is reasonable to
search for a group of patients presenting with typical paroxysmal vestibular and/or cochlear
symptoms, which are analogously caused by
neurovascular compression of the eighth cranial
nerve. If tinnitus is the major complaint, vascular
compression must cause abnormal impulse activity
of the cochlear portion of the eighth cranial nerve
(M0ller 1988; M0ller et al. 1992). If vertigo is the
major complaint, vascular compression must cause
abnormal impulse activity of the vestibular portion.
If vertigo and tinnitus are both present, then both
vestibular and cochlear portions of the eighth nerve
are compressed and partially demyelinated by one

or more vessels, either the anterior inferior cerebellar artery (AICA), the posterior inferior cerebellar
artery (PICA), or avein. Compression may be due to
vascular malformation, arterial ectasia of the posterior fossa (Yu et al. 1982; Buettner et al. 1983; Brandt
1991) (Fig. 7.4), or simply arte rial ageing with elongation and looping. It is well established that pulsatile compression of the caudal cranial nerves is
more likely to be symptomatic when the central
(oligodendroglia) rather than the peripheral myelin
is involved. For the eighth cranial nerve this means
that the entire intracranial portion from brain stern
to internal auditory canal (1-1.5 cm) is particularly
vulnerable.
Leigh and Zee (1991) reported hyperventilationevoked paroxysmal nystagmus and vertigo in single
cases of tumour compression of the eighth nerve at
the cerebellar pontine angle or the petrous bone. A
patient with Camurati-Engelmann disease, a progressive sclerotic bone dis order, was reported to
have fluctuating left-sided hearing loss, intermittent
high-pitched tinnitus, imbalance on rapid positional
changes and weekly episodes of spinning vertigo
lasting 5-10 minutes secondary to compression of
the eighth nerve in the left internal auditory meatus
(Fig. 7.5; Hanson and Parnes 1995). This patient was
successfully treated with internal auditory canal
decompression. The most likely mechanism involved
in the disorder was "neuro cross-talk" between focal
demyelinated axons.
The case of repetitive paroxysmal nystagmus and
vertigo described by Lawden et al. (1995) with complex combined torsional, horizontal and vertical nystagmus, however, is best explained by a central
vestibular nucleus lesion caused by an arteriovenous
malformation in dose proximity to the vestibular
nucleus. A brief burst of hyperactivity caused
episodic revers al and gross exacerbation of the resting nystagmus in this patient, who also responded
successfully to carbamazepine. The paroxysmal
change in the direction of the nystagmus reflects the
differential effects of hyper- and hypofunction of
vestibular structures (p. 124). The case of paroxysmal vertigo and spontaneous nystagmus evoked by
lateral eye position (Bttner et al. 1987) was also
central in origin (p. 242).

Management
Drugs: Carbamazepine (Tegretol) was administered
p.o. to all patients at an initial dosage of 3 x 100 mg
daily, increasing to a maximum of 2 x 400 mg daily
within 10 days. All patients responded promptly and

Peripheral vestibular paroxysmia (disabling positional vertigo)

123

a
Fig.7.5. A 28-year-old woman with Camurati-Engelmann disease, a progressive sclerotic bone disorder, which compressed the left
vestibular nerve and caused fluctuating hearing 1055, tinnitus and episodic vertigo. High-resolution thin section coronal computed
tomography of temporal bones at the level of the internal auditory canals (closed arrows). a Right b Left. Note slit-like opening of the
proximal half of the left canal. Also note diffuse sclerotic thickening ofthe temporal bone, particularly on the left superior surface (open
arrow). (From Hanson and Parnes 1995.)

significantly, even to the low initial dosage. Eight


became symptom-free; only three reported infrequent residual attacks. Efficacy of treatment has now
been followed for 6 years in one patient and for 2-3
years in five others who seem to require a minimal
dose of 200-400 mg per day. One or two washout
phases were performed in four patients, all of whom
relapsed within 2 days to 2 weeks. Carbamazepine
had to be replaced with phenytoin in two patients
due to allergie reactions. As phenytoin appeared less
effective, we tried pimozide (Orap). These two
patients are still taking it and have had complete
relief now for 2 and 3 years, respective1y (Brandt and
Dieterich 1994b).
Surgery: When the fifth, seventh and ninth nerves
are involved, microvascular decompression has
proved a well-established, successful treatment
(Jannetta 1982). For this reason, a few clinical centres
(Jannetta et al. 1984; M011er 1988; Ter Bruggen et al.
1987; M011er and M011er 1990; M011er et al. 1993)
have applied the same method for involvement of
the eighth nerve (carbamazepine was not tried in
this group of patients). They found vascular compression of the superior or inferior vestibular nerve
at its root entry zone. Microvascular decompression
gave relief from vertigo to most of these patients
(M011er et al. 1993). Benecke and Hitselberger (1988)
described a patient with a documented tortuous
basilar artery compressing the eighth nerve, in
whom selective section of the vestibular nerve provided complete relief from episodic vertigo. Hanson

and Parnes (1995) also described complete resolution of episodic vertigo after vestibular nerve
decompression in a case of Camurati-Engelmann
disease (Fig. 7.5).

Uncertainties in the diagnosis and treatment of


vestibular paroxysmia
The absence of a pathognomonic sign or test for
neurovascular compression of the eighth nerve
explains the critical reserve of some clinicians
(Brandt and Dieterich 1994b; McCabe and Harker
1983; Bergsneider and Becker 1995). Wiet and coworkers (1989) encountered a vascular loop in 6 of
36 patients who they considered suitable for retrolabyrinthine vestibular neurectomy for recurrent
vertigo. In view of the invasive procedure recommended in some centres once the diagnosis is established, prior circumspection is highly justified. The
surgical procedure, which involves retromastoid
craniotomy and microvascular decompression, is
still associated with a mortality of about 1% and a
considerable morbidity of about 10% (Modey 1985;
Friedmann et al. 1985), even though others report
lower mortality and morbidity rates. Delayed and
progressive hearing loss has been described as a rare
late complication, a result of reactive scar tissue and
progressive atrophy of the auditory nerve (Fuse and
M0ller 1996).

124

The hitherto proposed selection criteria for


surgery are insufficient, because they do not reliably
exclude causes of vertigo other than neurovascular
compression. Since the side with the lesion cannot
be identified solely by detecting vertigo,
audiovestibular tests and brainstem acoustically
evoked potentials (BAEP) are used to decide which
side is most probably affected and which one, therefore, should be explored surgically. BAEP recordings
with a unilaterally increased interpeak latency I-III
were reported to be the most helpful, since this indicates functionally significant lesions of the proximal
cochlear portion (peak II) of the eighth nerve. Such
recordings are similar to those seen in patients with
acoustic neuromas.
Preliminary reports from other clinical centres do
not help clarify the confusion of signs and symptoms seemingly characteristic of this enigmatic disease. All 10 patients diagnosed by Ter Bruggen and
co-workers (1987) as having neurovascular compression of the eighth cranial nerve were middle-aged
and hypertensive, most complained of paroxysmal
vertigo lasting for hours to days, which was aggravated by changing head position (movement), and
they had persistent tinnitus and progressive unilateral hearing loss. In contrast to Jannetta et al. (1984)
and M0ller et al. (1986), Ter Bruggen and co-workers
reported no positional vertigo, peak I of BAEP was
prolonged in six cases, whereas interpeak latency
I - III was normal in five cases, increased in one case
and undetected in three cases. Cerebrospinal fluid
total protein content was increased (0.6-1.5 gI-I) in
all patients; this unusual finding has not been
explained. Of the three patients who underwent surgical microvascular decompression, two improved
and one remained unchanged.
If selection for surgery is based on such
signs/symptoms, too many patients will be subjected
to an unnecessary and risky operation. The surprisingly high improvement rate (M0ller et al. 1986;
1993) is not adequate proof of correct diagnosis,
since other common vertigo syndromes such as
"phobic postural vertigo", can be treated rapidly and
successfully by suggestive or behavioural therapy in
more than 70% of patients (Brandt et al. 1994).
Furthermore, it is necessary to recall, for example, in
the management of Meniere's disease, that decades
passed before it was demonstrated that the apparent
success of the "endolymphatic sac shunt operation"
was actually due to the placebo effect (Thomson et
al. 1981). The varying success rates may be attributed
to both (1) the accuracy of differential diagnosis of
the underlying cause and (2) correct identification of
the affected side. It is not always possible to exclude
a central cause (e.g. central vestibular paroxysmia,
p. 243), especially if peripheral vestibulocochlear

Vertigo
deficits cannot be established in a patient presenting
with brief attacks of vertigo which respond to carbamazepine. Further progress in this disorder will
have to await advances in MRI technology that greatly
improve the visualisation of neurovascular crosscompression.

Alternating episodes of vestibular


nerve paroxysmia and failure
Vestibular nerve failure and excitation may occur in
alternating episodes in rare conditions of eighth

b
Fig.7.6. Compression ofthe eighth nerve by an arachnoid cyst.
a MRI, 3D-MP-RAGE, T,-weighted, axial. The vestibulocochlear
nerve (arrow) in the right cerebellopontine angle is displaced in
an anterior direction bya cyst. Apart of the cyst wall is delineated. b MRI, 3D-MP-RAGE, T,-weighted, coronal. The vestibulocochlear nerve (arrow) in the right cerebellopontine angle is
displaced in an upward direction by the cyst. (From Arbusow et
al. 1998.)

125

Peripheral vestibular paroxysmia (disabling positional vertigo)

R
L

t------l

10 mm

Fig. 7.7. Recordings of eye movements (top) and postural sway (bottom) during head rotations in a patient with eighth nerve compression by a cerebellopontine angle cyst (same patient as in Fig. 7.6). First fine: Electronystagmographic registration of eye movements under fixation and with closed eyes during an episode of vertigo. With the head in normal, upright position, spontaneous
nystagmus beating to the unaffected left ear (peak slow phase velocity 22/s) could be suppressed by visual fixation. Positioning of the
head to the right (45 around the vertical z-axis) had no distinct influence on spontaneous nystagmus; positioning of the head to the
left (45 around the vertical z-axis) transiently reversed spontaneous nystagmus, which now beats to the affected right ear. After 5 s
spontaneous nystagmus abruptly subsided without exponential decay. Second fine: Evaluation of body sway on a force-measuring
platform during an episode of vertigo. Measurement of the body sway with open eyes was within the normal range. With the eyes
closed and the head in anormal upright position, there was a slightly increased diagonal body sway from left-forward to right-backward. While positioning the head to the right did not substantially change body sway, positioning of the head to the left led to a distinct decrease of body sway together with a 90 shift in the preferred direction (from left-backward to right-forward). (From Arbusow
et al. 1998.)

nerve compression. Recurrent episodes of oscillopsia,


rotational vertigo and postural imbalance were
elicited and modulated by horizontal head positions
in a patient with an arachnoid cyst of the cerebellopontine angle (Fig. 7.6), wh ich displaced and
stretched the vestibulocochlear nerve (Arbusow et
al. 1998). Ocular motor analysis (Fig. 7.7) revealed
two different types of attacks depending on the particular head position: (1) episodes of vestibular
hypofunction (for minutes to several hours) with
normal head positions and (2) paroxysmal vestibular excitation (for seconds) with head rotation to the
left. Transition from conduction block to ectopic discharges, as occurs with varying peripheral nerve
compression (Ochoa and Torebj0rk 1980; Sugawara
et al. 1996), is the most likely pathomechanism. After
surgical resection of the cyst and decompression of
the eighth cranial nerve, vertigo, oscillopsia and postural imbalance completely resolved (Arbusow et al.
1998).

References
Andermann F, Cosgrove JBR, Lloyd-Smith D, Walters AM (1959)
Paroxysmal dysarthria and ataxia in multiple sclerosis.
Neurology 9: 211-215
Arbusow V, Strupp M, Dieterich M, Jger L, Hischa A, Schulz P,
Brandt Th (1998) Alternating episodes ofvestibular nerve excitat ion and failure. Neurology 51:1480-1483
Barroso L, Hoyt WF (1993) Episodic exotropia from lateral rectus
neuromyotonia - appearance and remission after radiation
therapy far a thalamic glioma. J Pediatr Ophthalmol Strabismus
30:56-57

Benecke JE, Hitselberger WE (1988) Vertigo caused by basilar


artery compression of the eighth nerve. Laryngoscope 98:
807-809

Bergsneider M, Becker DP (1995) Vascular compression syndrome


of the vestibular nerve: A critical analysis. Otolaryngol Head
Neck Surg 112:118-124
Brandt Th (1991) Vertigo: Its multisensory syndromes. London:
Springer
Brandt Th, Dieterich M (1990) Schwindel durch neurovaskulre
Kompression, "Vestibularisparoxysmie"? Nervenarzt 61:376-378

126
Brandt Th, Dieterich M (1994a) Vestibular paroxysmia: vascular
compression of the eighth nerve? Lancet I: 798-799
Brandt Th, Dieterich M (1994b) Vestibular paroxysmia (disabling
positional vertigo). Neuro-opthalmology 14:359-369
Brandt Th, Dieterich M (1 994c) Vestibular syndromes in the roll
plane: topographic diagnosis from brainstem to cortex. Ann
Neurology 36: 337-347
Brandt Th, Huppert D, Dieterich M (1994) Phobic postural vertigo: a first follow-up. J Neuro1241: 191-195
Buettner U, Sthr M, Koletzki E (1983) Brainstem auditory-evoked
potential abnormalities in vascular malformation of the posterior fossa. J Neuro1229: 247-254
Bttner U, Straube A, Brandt Th (1987) Paroxysmal spontaneous
nystagmus and vertigo evoked by lateral eye position.
Neurology 37: 1553-1555
Cohen B (1974) The vestibulo-ocular reflex arc. In: Kornhuber HH
(ed) Handbook of physiology, vol.V/2. Springer, Berlin,
pp 374-381
Dieterich M, Brandt Th (1993) Ocular torsion and tilt of subjective
visual vertical are sensitive brainstem signs. Ann Neurol 33:
292-299
Dieterich M, Brandt Th, Fries W (1989) Otolith function in man:
Results from a case of otolith Tullio phenomenon. Brain 112:
1377-1392
Friedmann W, Kaplan B, Gravenstein D, Rhoton A (1985) Intraoperative brainstem auditory-evoked potentials during posterior fossa microvascular decompression. J Neurosurg 62:552-557
Fuse T, Moller MB (1996) Delayed and progressive hearing loss
after microvascular decompression of cranial nerves. Ann Otol
Rhinol Laryngol 105: 158-161
Hanson W, Parnes LS (1995) Vestibular nerve compression in
Camurati-Engelmann disease. Ann Otol Rhinol Laryngol
104:823-825
Heimchen C, Dieterich M, Straube A, Bttner U (1992)
"Abduzensneuromyotonie" mit partieller Okulomotoriusparese.
Nervenarzt 63:625-629
Jacob RG, Moller MB, Turner SM, Wall C (1985) Otoneurological
examination in panic disorder and agoraphobia with panic
attacks: A pilot study. Am J Psychiatry 142: 715-719
Jger L, Strupp M, Brandt T, Reiser M (1997) Bildgebung von
Labyrinth und Nervus vestibularis. Nervenarzt 68:443-458
Jannetta PJ. (1975) Neurovascular cross-compression in patients
with hyperactive dysfunction symptoms of the eighth cranial
nerve. Surg Forum 26: 467-468
Jannetta PJ (1982) Treatment oftrigeminal neuralgia by microvascular decompression. In: Youmans J (ed) Neurological surgery,
vol6. Saunders, Philadelphia, pp 3589-3603
Jannetta PT, Moller MB, Moller AR (1984) Disabling positional
vertigo. N Engl J Med 310:1700-1705
Lawden MC, Phil D, Bronstein AM, Kennard C (1995) Repetitive
paroxysmal nystagmus and vertigo. Neurology 45:276-280
Leigh RJ, Zee DS (1991) The neurology of eye movements. Davis,
Philadelphia
Marks IM (1981) Space "phobia" : A pseudo-agoraphobic syndrome. J Neurol Neurosurg Psychiatry 44: 387-391
McCabe BF, Harker LA (1983) Vascular Jour as a cause of vertigo.
Ann Otol Rhinol Laryngol 92: 542-543
Moller AR (1991a) The cranial nerve vascular compression syndrome: I. A review of treatment. Acta Neurochir (Wien) 113:
18-23

Vertigo
Moller AR (1991b) The cranial nerve vascular compression syndrome: 11. A review of pathophysiology. Acta Neurochir (Wien)
113: 24-30
Moller AR, Moller MB, Jannetta PJ, Iho HD (1992) Compound
action potentials recorded from the exposed eighth nerve in
patients with intractable tinnitus. Laryngoscope 102:187-197
Moller MB (1988) Controversy in Meniere's disease: results of
microvascular decompression of the eighth nerve. Am J Otol 9:
60-63
Moller MB, Moller AR (1990) Vascular compression syndrome of
the eighth nerve. Neurol Clin 8: 421-439
Moller MB, Moller AR, Jannetta PT, Iho HD, Sekhar LN (1993)
Microvascular decompression of the eighth nerve in patients
with disabling positional vertigo: Selection criteria and operative results in 207 patients. Acta Neurochir (Wien) 125:75-82
Moller MB, Moller AR, Jannetta PJ, Sekhar LN (1986) Diagnosis
and surgical treatment of disabling positional vertigo. J
Neurosurg 64: 21-28
Morley TP (1985) Case against microvascular decompression in
the treatment of trigeminal neuralgia. Arch Neurol 42: 801-802
Nielsen VK (1984) Pathophysiology of hemifacial spasm: I.
Ephaptic transmission and ectopic excitation. Neurology
34:418-426
Ochoa JL, Torebjork HE (1980) Paraesthesiae from ectopic
impulse generation in human sensory nerves. Brain
103:835-853
Osterman PO, Westerberg CE (1975) Paroxysmal attacks in multiple sclerosis. Brain 98: 189-202
Rasminsky M (1980) Ephaptic transmission between single fibres
in the spinal root of dystrophic mice. J PhysioI305:151-169
Sanders DB (1989) Ephaptic transmission in hemifacial spasm: a
single-fibre EMG study. Muscle Nerve 12: 690-694
Sando I, Black FO, Hemenway WG (1972) Spatial distribution of
vestibular nerve in internal auditory canal. Ann Otol
81:305-314
Seltzer Z, Devor M (1979) Ephaptic transmission in chronically
damaged peripheral nerves. Neurology 29:1061-1064
Straube A, Bttner U, Brandt Th (1994) Recurrent attacks with
skew deviation, torsional nystagmus and contraction of the left
frontalis muscle. Neurology 44: 177-178
Sugawara 0, Atsuta Y, Iwahara T, Muramoto T, Watakabe M,
Takemitsu Y (1996) The effects of mechanical compression and
hypoxia on nerve root and dorsal root ganglia. Spine
18:2089-2094
Suzuki J-I, Tokumasu K, Goto K (1969) Eye movements from
single utricular nerve stimulation in the cat. Acta Otolaryngol
(Stockh) 68: 350-362
Ter Bruggen JP, Keunen RWM, Tijssen CC, Wijnalda D (1987)
Octavus nerve neurovascular compression syndrome. Eur
Neurol 27: 82-87
Thomson J, Bretlau P, Tos M, Johnson NJ (1981) Placebo effect in
surgery for Meniere's disease.Arch Otolaryngoll07: 271-277
Tomasulo R (1982) Aberrant conduction in human peripheral
nerve: Ephaptic transmission? Neurology 32:712-719
Wiet RJ, Schramm DR, Kazan RP (1989) The retrolabyrinthine
approach and vascular loop. Laryngoscope 99: 1035-1039
Yu Y, Moseley I, Pullicino P, Mc Donald W (1982) The clinical picture of ectasia of the intracerebral arteries. TNeurol Neurosurg
Psychiatry 45: 29-36

Bilateral vestibulopathy

Bilateral vestibular failure (BVF) is a rare disorder of


the peripherallabyrinth or the eighth nerve which
has various aetiologies. It is either acquired or congenital, or familial or sporadic. BVF occurs simultaneously or sequentially in both ears, and takes
either an abrupt or slowly progressive course. A
chronic bilateralloss of vestibular function is surprisingly weil tolerated. This is due to the fact that
spatial orientation, posture, and eye movements are
mediated by redundant multisensory processes
(p. 3), which can in part compensate for each other's
deficiencies. Moreover, there is no continuing distressing vertigo, pontaneous nystagmus, or postural
falls, wh ich are typical signs of a vestibular tone
imbalance caused by acute unilaterallesions. The
key symptoms are oscillopsia du ring locomotion or
head movements and unsteadiness, particularly in
the dark. The entity was first described by Dandy
(1941) in patients who had undergone bilateral
vestibular neurectomies. Generally patients with
BVF are first referred not only for assessment of
dizziness and disequilibrium, but also for examination of ocular motor disorders, ataxia, or hearing
loss, conditions in which BVF is often not suspected
prior to investigation (Rinne et al. 1995).
The diagnosis is made with the simple bedside
test for defective vestibulo-ocular reflex during
rapid, passive head turns (Fig. 8.1; Halmagyi and
Curthoys 1988; see also p. 39). It is confirmed by the
absence of nystagmic reachon to both caloric and
rotatory pendular testing while the patient sits in a
rotary chair. The most frequent aetiologies include
ototoxicity, autoimmune disorders, meningitis,
neuropathies, sequential vestibular neuritis, cerebellar degeneration, tumours, and miscellaneou otological diseases. So-called idiopathic BVF is found in
more than 20% of patients (Baloh et al. 1989; Vibert
et al. 1995; Rinne et al. 1995). Subjective symptoms in
the acute stage tend to improve with time by
processes of somatosensory and visual "substitution" of vestibular function. Vestibular rehabilitation
is supportive of this improvement, but the efficacy of

physical therapy is limited. The spontaneous recovery of patients with BVF is relatively rare and incomplete. A permanent loss of vestibular function is the
more frequent resuIt; however, the thus affticted
patient remains largely asymptomatic until confron ted with high-frequency motion conditions or
situations where proprioceptors or vision cannot
replace the deficient vestibular system (Brandt
1996).

The clinical syndrome


The two key symptoms of BVF are
unsteadiness o[ gait, particularly in the dark or
on unlevel ground
2. oscillopsia associated with head movements or
when walking.
1.

Episodes o[ vertigo lasting for minutes to days are


frequently reported by patients with sequential or
idiopathic BVF, particularly early in the development
of vestibular loss but not in the chronic state (Baloh
et al. 1989; Vibert et al. 1995; Rinne et al. 1995). Thus,
while episodic vertigo is compatible with the diagnosis of a newly acquired BVF, it is not a key symptom fo r this condition because of its diagnostic
ambiguity.
Unsteadiness in the dark or under impaired visual
conditions is the most frequent complaint. This is a
reflection of vestibulospinal deficiencies hindering
the multisensory process of postural contro!.
Somatosensory input from muscle spindies and
vision largely substitute for the missing vestibular
input. Therefore, patients feel stable when standing
or walking on Bat and firm ground under good visual
conditions. Imbalance occurs if either the visual or
the somatosensory input is reduced, as is the case

127

Vertigo

128

when the eyes are closed, in the dark, at high


altitudes (see Height vertigo, p. 418), when walking
on unlevel, soft, or wobbly ground, or in patients
with somatosensory polyneuropathy (see Nonvestibular vertigo, p. 447; Fig. 30.3). Patients may
report falls when going to the toilet at night in the
dark or when dancing in dimly lit rooms or their
inability to ride a bicycle at night or to ski in the fog
(Vibert et al. 1995). Imbalance also occurs with highfrequency head movements, because the detection of
high-frequency head rotation is a domain of the
vestibular apparatus. It has been shown that bilateral
labyrinthine-defective subjects are able to perform
linear goal-directed locomotion toward memorised
targets (Glasauer et al. 1994). Thus, the vestibular
system does not appear to be necessary for active
linear path integration.
Oscillopsia, the illusory movement of the viewed
stationary scene, is associated with head movements,
when walking, or during locomotion by vehicle. It
reflects an insufficiency of the vestibulo-ocular
reflex at higher head rotation frequencies. The
vestibulo-ocular reflex normally holds the direction
of gaze in space constant during head movements,
by driving the eyes to move in their orbits in the
direction opposite that of head motion at a velo city
and amplitude which "compensates" for the head
motion. If the amplitude and/or velo city of eye
movements are inappropriate, there is a shift in the
direction of gaze. This causes a displacement or slip
of the retinal image, which may be perceived as an
apparent motion of the fixated object (p. 430).
Oscillopsia is always less than retinal image slip in
deficient vestibulo-ocular reflex due to the impaired
perception of motion characteristic of this condition
(Brandt and Dieterich 1988; Fig. 29.20, p.432).
Therefore, it is not surprising that only 30-40% of
patients with BVF spontaneously complain about
oscillopsia (Baloh et al. 1984; Vibert et al. 1995).
Oscillopsia occurs most frequently in young
patients, at the onset of the disease (Bhansali et al.
1993), and in severe (complete) vestibular loss
(McGarth et al. 1989; Telian et al. 1991). Oscillopsia
differs from perception of real motion in that it creates a disturbing experience of spatial disorientation
and impairs the ability to recognise faces or to read
while walking. A small subgroup of patients may
even experience oscillopsia when the head is kept
still, e.g. when reading or watching television.
Bronstein et al. (1992) labeled this combination of
BVF and essential head tremor "pendular pseudonystagmus". The involuntary ocular oscillations
observed are in this condition caused by the head
tremor and the insufficient vestibulo-ocular reflex.
Oscillations of the fundus can be detected by

ophthalmoscopy (Zee 1978) and are attenuated or


suppressed in these patients by rigid immobilisation
of the head (Bronstein et al. 1992).
The course of BVF, as it involves the right and left
labyrinths (or eighth nerves), can be
1. simultaneous or
2. sequential,

and either
1. abrupt or
2. slowly progressive.
The severity of BVF can be
1. incomplete (moderate or severe bilateral vestibu-

lar failure) or
2. complete (bilateral vestibular loss).

Diagnosis
The suspected diagnosis can be supported by a useful bedside test of the vestibulo-ocular reflex (Fig.
8.1). If the head of the patient is turned passively and
quickly (say, faster than 1 Hz), then the compensatory
eye movement needed to maintain gaze in space is
mediated by the vestibular system rather than the
optokinetic-pursuit reflex or cervico-ocular reflex.
In bilateral vestibulopathy - despite attempted fixation of a stationary target - the gaze shifts with the
head (because compensatory eye movements are
inappropriate). After the head movement, a compensatory saccade toward the fixation target corrects the
gaze in space, a process that can be easily observed
(Halmagyi and Curthoys 1988). These tests do not
exclude the possibility that parts of the vestibular
labyrinth may still function, especially the vertical
semicircular canals or the otoliths.
The diagnosis can be confirmed by bithermal
caloric testing (if necessary, using the more intense
stimulus of iced water irrigation) and pendular body
rotation on a rotary chair in the dark, e.g. at a frequency of 0.05 Hz and high velocities of 80 or 100 0 /s.
A complete bilateral vestibular loss is defined by the
absence of both bilateral caloric responses and
vestibular rotation al responses (gain = peak slow
phase eye velo city/peak chair velo city < 10% of
mean normal value; Rinne et al. 1995). Moderate and
severe BVF are defined less consistently according to
the quantitative diminishment of the caloric and
rotational responses. Caloric irrigation is the more
informative test, because it is the stronger stimulus
(iced water) and allows better differentiation of the
remaining right and left ear minimal function.

129

Bilateral vestibulopathy

Fig.8.1. Vestibulo-ocular reflex bedside test. Top. Norma l gaze fixation du ring rapid head turn toward intact side. a,b With her face
turned a little to the right and her eyes fixed on a distant target, the patient (professional model) waits for her head to be moved
rapidly to the left by the examiner. C After leftward head movement, gaze is still fixed on the target so that no refixation saccades are
required.Bottom. Clinical signs of right semicircular canal paresis. Abnormal gaze fixation during rapid head turn toward the lesioned
side. a With her face turned a little to the left and with her eyes fixed on a distant target, the patient (professional model) waits for
her head to be moved rapidly to the right. b Following rightward head turn, it becomes evident that the gaze has shifted during
head turn with head to the right. c Leftward or compensatory saccade is now required to refix the gaze (From Halmagyi and
Curthoys 1988.)

Rotational testing is also very useful for eonfirming


diagnosis, espeeially in rare eases where thermal
irrigation is eontraindieated, e.g. tympanie perforation.

Associated symptoms and differential


diagnosis
The differential diagnosis of BVF is approaehed
from two direetions. On the one hand, the various
pathologies eausing the loss of vestibular funetion
are differentiated. On the other, BVF is differentiated
from other vertigo syndromes, whieh also eause
unsteadiness and oseillopsia.
Assoeiated symptoms are espeeially helpful for
the differential diagnoses of the various pathologies
underlying BVF. Associated hearing loss in BVF is
eommonly due to otologieal diseases, neuropathies,
meningitis, and autoimmune disorders. It is less
eommonly eaused by ototoxicity or eerebellar
degeneration, and it seldom oeeurs in the idiopathie
type of this eondition (Rinne et al. 1995). The eombination of BVF with hearing loss and signs of
multisystemie disease (including inflammatory eye

symptoms) is indieative of an auto immune disease


(Table 8.1), whereas associated ataxia or sensory loss
points to eerebellar degeneration or neuropathy
(Rinne et al. 1995). CT and MRI seans and eerebrospinal fluid (CSF) examination are helpful for
identifying rarer eauses of BVF (see Tables 8.1-8.3).
The differential diagnoses most relevant to BVF
are eerebellar or oeular motor dis orders without
BVF, intoxieation (p. 393), phobie postural vertigo
(p. 469), vestibular paroxysmia (p. 117), (otolithie)
perilymph fistula (p. 99), eervieal vertigo (p. 441),
orthostatie hypotension, hyperventilation syndromes,
visual disorders, and unilateral vestibular loss
(p.71).

Aetiologies and pathomechanisms


A rare eondition, BVF aecounts for 2.6% of the diagnoses in our dizziness unit (52 of 2010 patients;
Table 2.1). The age of patients with BVF ranges from
youth to old age, with a mean age at diagnosis in the
fifties (52.4 19.1 years; Telian et al. 1991). The age

Vertigo

130
Table 8.1. Diagnosis of immune-mediated inner ear disease
Clinical presentation
Rapidly progressive over days, weeks, or months
Usually bilateral, but onset asymmetrieal; may be unilateral
Vestibular symptoms variably present
Pressure and tinnitus ohen present
More eommon in middle-aged women but may oeeur in both sexes at
anyage
Presenee of systemie disease should be considered
Audiological examination does not identify a eause
MRI or Cl: no lesion or demyelinating disease
AER usually eonsistent with peripheral disease
Ophthalmie examination
Laboratory examinations that ean be useful but whieh are not specifie for
immune-mediated inner ear disease
Complete blood count
Erythroeyte sedimentation rate
FTA-ABS
Immune complex sereen
Antinuclear antibody
If history suggests ean add
Sinus and ehest radiography
Antineutrophil eytoplasmie antibody
Anti-Borrelia burgdorferi antibody
Rheumatoid faetor
Laboratory examinations specifie for immune-mediated inner ear disease
Lymphoeyte proliferation (must be done before treatment)
Western blot immunoassay for antibody to 68-kDa inner ear antigen
(available soon)
Trial of immunosuppression (eortieosteroids, eyclophosphamide)
AER, auditory evoked response; FTA-ABS, fluoreseent treponemal
absorption assay.
From Moscieki (1994).

of manifestation obviously depends on the aetiology.


No relevant sexual predominance is noticeable. A
review of several thousands of unselected
electronystagmograms in different neurological and
ENT settings revealed that BVF has an overall frequency of 0.6-2% (Simmons 1993; McGarth et al.
1989; Vibert et al. 1995). Both the kind and the frequency of the particular dis order causing BVF
depend considerably on whether the statistics are
presented by neurologists or otolaryngologists
(Table 8.2a,b). Whereas ototoxicity is a major ENT
diagnosis, cerebellar degeneration or neuropathies
are most frequently diagnosed in neurology.
Ototoxie BVF is mostly due to antibiotic treatment
with gentamicin (Schuknecht 1974) alone or in combination with erythromycin (Agusti et al. 1991) or
other antibiotics. ld age and renal failure are wellrecognised risk factors for increased sensitivity to
ototoxicity (Barza and Lauermann 1978).
Streptomycin and gentamicin are known to affect
the peripheral vestibular sensory cells before those
in the cochlea (hair cell damage of the inner ear).
Kanamycin, tobramycin, and neomycin preferentially
impair the auditory sensory cells (Schuknecht 1974),

Table 8.2a.

An ENT diagnosis of patients showing bilateral vestibular


loss at the neuro-otologieal examination (n = 52); 14 (28%) of the 25
"idiopathie" patients presented with normal hearing and without neurologieal symptoms

Idiopathie
Ototoxieity
(Garamycine)
Brainstem tumours
(bilateral aeoustie neurinoma, lymphoma metastasis)
Heredo-degenerative diseases
(Usher, Friedreieh, Charcot-Marie)
Meningitis
Bilateral labyrinthitis
Bilateral temporal bone fraeture
Bilateral Meniere's disease
AIDS
Bilateral inner ear fistula
(eholesteatoma)

25 (48%)
11

2
5
2
2
2
1

From Vibert et al. (1995).

Table 8.2b.

A neurological diagnosis of 53 patients with bilateral


vestibular failure (BVF)
BVF in association with neurologieal disease
Cerebellar degeneration
Neuropathies
Miseellneous
Idiopathie BVF
Ototoxicity
Post -meningitie
Autoimmune
Miseellaneous otologieal diseases
Neoplastie (acoustie neuromas excluded)

15 (28%)
7(13%)
5(9%)
3(6%)
11 (21%)
9(17%)
6(11%)
5 (9%)
4(7%)
3(6%)

From Rinne et al. (1995).

whereas more recent aminoglycosides such as


dibekacin or ribostamycin are less ototoxic (Sato
1982). Hair cells and cochlear neurons may be
transiently (reversibly) damaged by diuretics
(furosemide) or high-dose salicylate therapy (Myers
et al. 1965; Schuknecht 1974). Combined use ofloopinhibiting diuretics and aminoglycoside antibiotics
can cause permanent hearing loss (Baloh 1984).
Permanent loss of hair cells sometimes occurs with
alkylating anti cancer chemotherapeutics or cisplatin
(Schuknecht 1974) (see also Chap. 28, p. 395).
Cerebellar degeneration, sporadic multisystem
degeneration, or Friedreich's ataxia can also cause
BVF. Cerebellar ataxia, dysarthria, clumsiness of the
hands, or ocular motor abnormalities should prompt
neurological investigation. Accordingly, BVF is often
an unsuspected finding in routine electronystagmography (Rinne et al. 1995; Vibert et al. 1995).
Unsteadiness in the dark and oscillopsia in these
patients can be explained by both dis orders, cerebellar degeneration and BVF. The site of degeneration,

Bilateral vestibulopathy

131

peripheral and/or central, has not yet been systematically investigated with histopathological methods.
In 9% of aseries of patients (Rinne et al. 1995)
BVF was associated with cranial or peripheral
neuropathies, such as vitamin B12 deficiency, sarcoidosis (Jahrsdoerfer et al. 1981), alcohol (Ylikoski
et al. 1981), hereditary sensory and autonomie
neuropathy type IV (HSAN IV) (Pinsky and
DiGeorge 1966), and nutrition al (beri beri) neuropathy (Gill and Bell 1982). A striking synchronisation of the c1inical course inc1uding relapses and
responses to immune therapy was reported to occur
in patients with chronic inflammatory demyelinating polyneuropathy and bilateral vestibulopathy
(Frohman et al. 1996).
The post-meningitic group accounted for 11 % of
the cases of BVF in the study of Rinne et al. (1995);
three cases were attributed to Streptococcus suis
meningitis through contact with infected pigs
(Lamont et al. 1980). Streptococcus pneumoniae,
Neisseria meningitidis, Mycobacterium tuberculosis,
Borrelia burgdorferi (Lyme disease, p. 151), and
probably also the human immunodeficiency virus
(HIV) (Husler et al. 1991; Grimaldi et al. 1993;
Vibert et al. 1995) are other causative agents. The frequency of HIV-BVF is likely to increase with time.
Systemic autoimmune diseases can cause rapidly
Fig.8.2. Cogan's syndrome, subacute stage with bilateral hearprogressive, usually bilateral (although the onset is ing 1055 and vestibular loss.a MRI axial projection.T,-weighted
often asymmetrical) sensorineural hearing loss and 20 FLASH. Hyperintense lesions, a sign of subacute haemorBVF (Hughes et al. 1984; Moscicki 1994, Table 8.1). rhage, are detected in the vestibule (short arrow) and in the
Cogan's syndrome is the protypical immune disease cochlea (long arrow). b MRI axial projection, T,-weighted 20
FLASH, post Gd-OTPA. The same haemorrhagic lesions can be
affecting the inner ear (Fig. 8.2) (Cogan 1945; Cody delineated
by an enhancement of the cochlea (long arrow) and
and Williams 1960; Bicknell and Holland 1978; the vestibule (short arrow). (From Jger et al. 1997.)
Haynes et al. 1980; McDonald et al. 1985; Vollertsen
et al. 1986; Terjung et al. 1993; Cote et al. 1993; see
Sequential
bilateral
vestibular
neuritis
also p. 154). Auditory and vestibular failure has also
been reported in Beh<;:et's disease (Barna and Hughes (Schuknecht and Witt 1985; Ogata et al. 1993)
1988; Tsunoda et al. 1994) and sarcoidosis accounts for an unknown percentage of BVF, prob(Jahrsdoerfer et al. 1981). Other immune-mediated ably also for apart of the so-called idiopathic BVF.
systemic diseases associated with auditory and Bilateral Meniere's disease (p. 86) may also cause
vestibular failure (Moscicki 1994) are systemic lupus moderate or severe BVF and some authors stress the
erythematosus, polychondritis, juvenile rheumatoid link between bilateral Meniere's disease and
arthritis, adult rheumatoid arthritis, cerebral vas- immune-mediated mechanisms (Suzuki and Kiahara
culitis, polyarteritis nodosa (Rowe-Jones et al. 1990), 1992).
BVF is one of the rare causes of vertigo, but also
Wegener's granulomatosis, giant cell arteritis,
Sjgren's syndrome, and possibly elevated antiphos- one with the most different aetiologies. Numerous
pholipid antibodies (Rinne et al. 1995), which may rare causes have been reported (Table 8.3). Alport's
cause recurrent venous thrombosis (Harris et al. syndrome (an inherited sensorineural deafness asso1988). Serum antibodies against membranous ciated with interstitial nephritis), Waardenburg's
labyrinth have also been found in patients with syndrome (an inherited deafness associated with
"idiopathic" BVF (Arbusow et al. 1998).
facial dysplasia), bilateral Mondini dysplasia
Neoplastic causes of BVF are bilateral acoustic (Schuknecht 1980; Kommune et al. 1993), and Usher's
neuromas (Young et al. 1970) in neurofibromatosis syndrome (an inherited sensorineural deafness asso(Type II) (Fig. 8.3), non-Hodgkin's lymphoma, lep- ciated with retinitis pigmentosa, Trop et al. 1995)
tomeningeal metastasis (Fig. 8.4), and infiltration of may all be accompanied by bilateralloss of vestibuthe skull base.
lar function (Nance and Sweeney 1975; Knigsmark

132

Vertigo

Fig.8.3. MRI of a patient suffering from neurofibromatosis (type NFII). a Axial and b frontal MRI orientations after application of
intravenous Gd-DTPA (TR/TE = 500/17 ms) show multiple neuromas involving cranial nerves bilaterally. On the right side, there is an
intracanalicular acoustic neuroma as weil as a hypoglossal neuroma. On the left side, an extracanalicular neuroma has developed 3
years after surgical removal of an acoustic neuroma; it is pressing on and dislodging the pontine brainstem.

and Gorlin 1976; Baloh 1984; Wester et al. 1995;


Huygen and Verhagen 1994). Genetic diagnosis is
possible in most of these dis orders (See also Chap.
26, p. 378). A familial vestibular areflexia with mild
oscillopsia and imbalance but without hearing loss
was documented by Verhagen et al. (1987) in three
siblings; it was presumed to arise from autosomal
recessive inheritance. Baloh et al. (1994) described a
dominantly inherited bilateral familial vestibulopathy in three patients who presented with episodic
vertigo followed by gait imbalance and oscillopsia
with bilateral vestibular loss despite normal hearing.
Single cases of"progressive vestibular degeneration"
of unknown aetiology have been described earlier,
but all had the following factors in common: repeated episodes of dizziness relatively early in life, bilateral loss of vestibular function with retention of
hearing, and freedom from other neurological disturbances (Diamant 1946; Guttich and Stark 1965;
Graybiel et al. 1972). Atypical Meniere's disease was
ruled out in these cases by the absence of cochlear
symptoms and the time course of the condition.
Other rare causes or single case descriptions
relate BVF to bilateral occipital or temporal bone
fracture (Feneley and Murthy 1994), labyrinthitis
fibro-ossificans (Lu and Schuknecht 1994), mitochondriopathy (Lenard et al. 1992), macroglobulinaemia (p.288), vertebrobasilar dolichoectasia
(Bttner et al. 1995; Nuti et al. 1996) (Fig. 8.5).

Vestibular atelectasis (Merchant and Schuknecht


1988) is another possible mechanism for unilateral
or bilateral vestibular dysfunction: the walls of the
ampulla and utricle simply collapse. New bone formation or excessive fibrous tissue can damage
peripheral auditory and vestibular structures in
Paget's disease (Schuknecht 1974) or CamuratiEngelmann disease (Hanson and Parnes 1995) (Fig.
7.5, p. 123). Otosclerosis normally presents as a hearing dis order that results from immobilisation of the
stapes. In ab out 25% of these patients it is also associated with unsteadiness of gait or episodic vertigo
(Thomas and Cody 1981). It is probably caused by
otosclerotic foci that affect the vestibular nerve or
sensory cells. There is reason to assurne that some
cases of BVF may have a vascular cause. It has been
suggested that the superior vestibular artery - which
supplies the horizontal semicircular canal and the
utricular macula (Mazzoni 1974) - is especially vulnerable to ischaemia because of the small calibre of
the anterior vestibular artery and the lack of collaterals (Grad and Baloh 1989).

Idiopathic BVF

The underlying cause can be identified in about


70-80% of patients presenting with BVF. However,
more than 20% of such patients have idiopathic BVF
with normal hearing and no associated neurological

133

Bilateral vestibulopathy

b
Fig.8.4. Meningeal carcinomatosis of a lung cancer with metastases in the labyrinth.a MRI axial projection, T2 -weighted 3D-ClSS.
Recesses in the cerebellopontine angle (black arrow), the internal acoustic canal, and the cochlea (white arrow) indicate metastases.
b MRI axial projection, T,-weighted 2D-FLASH, post Gd-DTPA. Metastases are shown by enhancement in the internal acoustic canal (/arge
arrow), cochlea (smalliong arrowL vestibule (arrowhead), and the lateral semicircular canal (5mallshort arrow). (From Jger et al. 1997.)

symptoms (21 %, Rinne et al. 1995; 27%, Vibert et al.


1995). The mean age of the reported patients varies
from 38 years (Vibert et al. 1995) to 57 years (Baloh
et al. 1989), with a range of 20 to 84 years; the mean
age of onset may be earlier. About 50% of this group
report prior episodes of vertigo (Baloh et al. 1989) or

episodes of head movement-induced oscillopsia


with unsteadiness or brief attacks of spontaneous
vertical oscillopsia lasting for seconds (Rinne et al.
1995). Idiopathic BVF cannot be explained simply by
vestibular ageing, although the peripheral vestibular
system does show a progressive loss of sensory cells

134

Vertigo

Table 8.3. Frequent and rare diagnoses in bilateral vestibular failure


(BVF)
Relatively frequent
-Idiopathic (20-30%)
- Ototoxicity
- Cerebellar degeneration
- Meningitis (labyrinthine
infections)

- Tumours

- Autoimmune disorders

- Neuropathies

- Bilateral sequential
vestibular neuritis
- Bilateral Meniere's disease

Gentamicin and other antibiotics


Anticancer chemotherapy
Sporadic multisystem degeneration
Friedreich's ataxia
Streptococcus suis
Neisseria meningitidis
Mycobacterium tubercu/osis
Streptococcus pneumoniae
Borre/ia burgdorferi
Human immunodeficiency virus (HIV)
Neurofibromatosis (bilateral acoustic
neuromas)
Non-Hodgkin's lymphoma
Leptomeningeal metastasis
Infiltration of skull base
Cogan's syndrome
Beh~et's disease
Cerebra I vasculitis
Systemic lupus erythematosus
Polychondritis
Rheumatoid arthritis
Polyarteritis nodosa
Wegener's granulomatosis
Giant cell arteritis
Primary antiphospholipid syndrome
B12 defieiency
Hereditary sensory and autonomie
neuropathy
(HSMN IV)
Nutritional (beri beri)
Neurosarcoidosis

Delayed endolymphatic hydrops

Relatively rare (or single cases)


- Congenital malformation
-

Fig.8.5. Bilateral vestibular 1055 in vertebrobasilar dolichoectasia. The ectatic and tortuous basilar artery extending to the left
side and apparently compressing the brainstem can be clearly
seen. Note that there is no sign of bleeding or infarction. T,weighted axial MRI scan (TR = 570 ms, TE = 15 ms) at the pontine
level after contrast enhancement (Gd-DTPA) (From Bttner et al.
1995.)

Usher's syndrome and other rare


conditions
(see Chap. 26, p. 378)
Familial vestibulopathy (progressive vestibular degeneration)
Bilateral temporal bone fraeture
Vertebrobasilar doliehoeetasia
Camurati-Engelmann disease
Paget's disease
Vestibular ageing (?)
Vertebrobasilar ischaemia
Labyrinthitis fibro-ossifieans
Vestibular ateleetasis
Mitochondriopathy
Macroglobulinaemia

and afferent neurons with age (Johnson and


Hawkins 1972), an effect similar to presbyacusis.
This loss does not, however, have a severe effect on
the vestibulo-ocular reflex. Age-dependent deterioration was most pronounced in phase measures
(phase lead increases with ageing), both at low
frequency and low head velo city, and at modest

frequency but high head velo city (Paige 1992). Gain


decrements were also observed with ageing, but the
changes were more subtle. Thus, ageing entails a
slowly progressive bilateral vestibular failure, but
ageing alone should not be confused with the "disorder of idiopathic BVF" or the frequent falls in the
elderly (see Chap. 27). Here "idiopathic" stands for
unknown aetiology; however, some of these cases
may be due to undetected auto immune disorders or
ischaemia of the labyrinth.
To investigate the possibility of an auto immune
mechanism in idiopathic BVF patients, Arbusow and
co-workers (1998) screened sera for antibodies
against inner ear structures. They found IgG antibodies against membranous labyrinth (ampulla,
semicircular canals, saccule, utricle) in 8 of 12
patients by using immunofluorescence on rat inner
ear cryosections (Fig. 8.6). While antilabyrinthine
autoantibodies may be an epiphenomenon of BVF, a
small subgroup of organ-specific autoantibodies
could synergise with a cellular response during the
development of vestibular lesions.

Bilateral vestibulopathy

13S

Fig.8.6. Immunofluorescent labelling of rat inner ear cryosections by serum IgG (diluted 1:100) from a patient with "idiopathic"
bilateral vestibular failure and contral serum. a Ampulla, patient serum; b ampullary control serum; c semicircular canal, patient serum;
d semicircular canal, contral serum; e utricle patient serum; f ampulla, patient serum, double staining with TRITC-phalloidin. Vestibular
hair cells are stained red. (Fram Arbusow et al. 1998.)

Spatial orientation: vestibulo-ocular and


vestibulospinal reflexes
A failure of vestibular function will produce as a
consequence signs and symptoms of dysfunction in
all three major sensorimotor processes comprising
the vestibular system - spatial orientation, stabilisation of gaze in space, and postural control (Fig. 8.7,

p. 137). For example, spatial orientation in BVF


patients should suffer from a diminished ability to
perceive verticality and self-motion (velo city and
change of head position in space), particularly under
stimulus conditions with reduced and thus less reliable input from other sensory sources, such as vision
or somatosensors. These patients may experience
disorientation when diving or riding in vehicles;

136

their reduced sense and control of dynamic rotational


orientation has been experimentally demonstrated.
Similarly, labyrinthine-defective subjects failed to
sense, to any useful extent, their rotation on a rotary
chair in the dark (Brookes et al. 1993). Patients with
bilateral vestibular loss are subject to stronger
effects (and disorientation) of somatosensory
(p. 447) and optokinetic stimulation (Bles et al.
1983). When objectively stationary patients are
exposed to a large visual display rotating ab out an
earth-horizontal axis, they experience a complete,
apparent self-rotation through an upside-down
orientation (Cheung et al. 1989). In contrast, a
healthy stationary observer, while viewing a large
visual scene rotating around his line of sight, experiences a continuous sensation of self-motion opposite
in direction to pattern motion (rollvection), but only
a limited body tilt and displacement of the visual
vertical (Dichgans et al. 1972). Induced displacement
can be conceptualised as the result of a compromise
between the different and, in part, contradictory sensory inputs for gravitational orientation (p.414).
Patients with BVF have a persisting directionspecific (horizontal> vertical) reduced sensitivity
for detecting visual motion of single objects, even
when the head and eyes remain still (Grnbauer et
al. 1998).
The absence of the vestibulo-ocular reflex is
largely substituted for by visual/optokinetic
(Dichgans and Brandt 1978) and somatosensory
input, especially by the cervico-ocular reflex.
Whereas the gain of the cervico-ocular reflex is low
in normal subjects, its sensorimotor weight is significantly enhanced in patients with vestibular loss
(Kasai and Zee 1978; Chambers et al. 1985; Bronstein
and Hood 1986). Visual information, however, can
greatly modify the gain of the cervico-ocular reflex
in these patients (Heimbrand et al. 1996). An
increase in gain following an acute vestibular deafferentation requires weeks to become functionally
significant, a finding that agrees with experimental
data in the labyrinthectomised monkey (Dichgans et
al. 1973). The reversibility of such changes in gain
was demonstrated in a patient with post-meningitic
vestibular loss who gradually recovered within
months (Bronstein et al. 1995). Along with the
patient's vestibular recovery and increasing vestibuloocular reflex gain, the compensatory high cervical
gain decreased toward its normal level. Using a
"remembered saccade" technique, Nakamura and
Bronstein (1995) found there was no significant
change in the ability of labyrinthine-defective
patients to detect head turns (head to trunk).
Rotational testing in patients (Baloh et al. 1984;
Mller and dkvist 1989; Bhmer and Fisch 1993)
and monkeys (Waespe et al. 1992) demonstrated the

Vertigo

frequency-dependent gain and the abolished "velocity storage" mechanism. In the dark patients have no
or very low gains, but when tested with a stationary
light or even with imaginary stationary targets in the
dark their gain can be enhanced to normal values
with a phase lead in the low-frequency range; even a
stationary acoustic source can enhance ocular motor
gain (Mller and dkvist 1989). The significant
compensation for a vestibular deficit in the pursuit
of moving targets with combined eye and head
movements was demonstrated by Waterston et al.
(1992) for slow-phase gaze velocity gains under
head-free and head-fixed conditions. They found no
significant difference from pursuit in normal subjects. Somatosensory substitution for vestibular
function is not only measurable by the increased
gain of cervico-ocular reflex but also in arthrokinetic
nystagmus and the sensation of self-motion (p. 446)
as induced by limb movements. Patients with bilateral
loss of labyrinthine function exhibit characteristic
abnormalities of arthrokinetic nystagmus: drastic
shortening of latencies, rapid build-up of slow phase
velo city, increased gain and absence of afternystagmus (Bles et al. 1983). Further substitution of
vestibular function is achieved by refixation saccades and behavioural strategies such as restriction
of head movements.
Vestibulospinal consequences of bilateral vestibular loss have been investigated in animal
experiments and patients (Fig. 8.7). Bilaterally
labyrinthectomised cats are able to stand unsupported
on a force platform with little changes in stance
parameters from those of the control, pre-lesion
state (Thomson et al. 1991). There was no change in
the distribution of vertical forces under the limbs
and no increase in sway; the horizontal plane forces,
which had a diagonal direction prior to lesion, took
a more lateral direction and became larger in amplitude. Postural responses of the cats were characterised by normal latency and normal spatial and
temporal patterning of electromyographic response
(Ingles and Macpherson 1995). The only deficit in
the postural response after lesion was a hypermetria
or overactive response that caused the animal to
overbalance somewhat, but did not impair its ability
to remain upright. It was concluded "that vestibular
information is not essential for triggering the rapid,
automatic postural response to translations of the
support surveys" (however, it may trigger rapid postural responses; see p. 108), "nor is it necessary for
the selection or shaping of the evoked response.
Instead, somatosensory information appears to predominate in these postural adjustments. However,
vestibular afferent input does influence the scaling of
the postural response" (Ingles and Macpherson
1995). Accordingly, children with congenital or early

137

Bilateral vestibulopathy

EYES OPEN

normal

EYES CLOSED

R---+----L

10mm
t-----I

P
Bilateral Vestibulopathy

displacement were absent in patients who had lost


vestibular function (Allum et a1. 1994). Allum and
co-workers demonstrated that bilateral vestibular
deficit subjects respond to balance perturbations
under eye-open conditions, as if the perturbations
were 50% slower. The perturbation of choiee when
testing vestibular deficit patients was a rotation
rather than a translation of the support surface.
Finally, Herdman et a1. (1994) found increased body
sway on sensory tests in which either visual or
somatosensory cues were alte red, and persistent falls
when both cues were altered. The latter is best
demonstrated by the patient suffering from a combination of BVF and sensory polyneuropathy (Fig.
30.3, p. 448), who under scotopic visual conditions
was unable to maintain balance (Paulus et a1. 1987).

Management

K.J.<jl50

Fig. 8.7.

Postural instability in bilateral vestibulopathy.


Histograms for fore-aft (A,P) and lateral (R,L) postural sway during
upright stance with eyes open (feft) and eyes closed (right)
obtained with a force-measuring platform. For comparison see
registration of body sway in anormal subject (top). Preferred
direction of postural instability and body sway is in the fore-aft
direction for bilateral vestibulopathy (bottom). Body sway increases significantly with eyes closed when visual stabilisation cannot
substitute for the peripheral vestibular deficit. Force-plates do not
measure body displacement, instead they measure the position
of the centre offoot pressure (CFP). The movement of this point is
influenced both by displacement of the centre of gravity and by
body acceleration (produced by muscle activity).lf body displacements are slow, the CFP will indicate approximately the position
of the centre of gravity, but with active postural correction CFP
displacements can be considerably larger. The histograms shown
here probably contain both components; however, this does not
alter the clinical interpretation much (From Brandt and Dieterich
1996).

acquired bilateral vestibular loss do not show an


increase in body sway with respect to "normal" children while standing on a bare surface (Enbom et a1.
1991). Body sway velocities were found to increase
more in children with BVF than in normal children
when both were standing on foam rubber. Patients
with bilateral vestibular loss showed patterns and
latencies of leg and trunk muscle responses to body
displacements similar to those of healthy subjects
(Horak et a1. 1994). In contrast to responses to body
displacements, responses to direct head displacements appeared to depend on a vestibulospinal trigger, since trunk and leg muscle responses to head

Management of the various dis orders causing BVF


has three aims:
1. prophylaxis of progressive vestibular loss

2. recovery of vestibular function

3. facilitation of compensation (substitution) for


vestibular function by physical therapy.
Prevention is most important for the group of
patients with ototoxic BVF. Careful monitoring of
audiograms and caloric testing is required with
aminoglycoside therapy. This is, however, not possible in the critically ill patient. Magnusson et a1.
(1991) reported that the ototoxic effects of gentamicin have a delayed onset. This has very important
consequences:
a) an extremely low dose of gentamiein may suffice
to produce an adequate lesion of the vestibular
function, and
b) there may be a danger of the drug having excessive ototoxie effects.
Therefore, the usual strategy of applying aminoglycosides until signs of ototoxie effects appear is
more dangerous than previously imagined (Brandt
and Strupp 1992). Recognised risk factors for ototoxicity are renal failure and old age (Barza and
Lauermann 1978), and even a familial susceptibility
to aminoglycoside ototoxicity has been described
(Hu et a1. 1991).
Prevention and recovery are possible by treating
immune-mediated inner ear disease. The treatment
should be initiated with corticosteroids as early as

Vertigo

138

possible, e.g. prednisone in doses of 60 mg daily for Table 8.5. Bilateral vestibular failure (BVF)
adults and 1 mg/kg body weight for children daily
Clinieal syndrome
for aperiod of at least 30 days (Moscicki 1994).
Symptoms
Improvement may not be symmetrical, usuaHy the
- Unsteadiness of gait (partieularly in the dark or on unlevel ground)
most recently affected ear responds best. If the
- Oscillopsia associated with head movements or when walking
- Episodes of vertigo early in the development of BVF but not in
response is inadequate, addition of cydophoschronic state
phamide or azathioprine or methotrexate may be
considered as weH as plasmapheresis or high-dose
Signs
- Pathological vestibulo-ocular reflex-bedside test
intravenous gammaglobulins (for treatment of
- Absent vestibulo-ocular reflex with bithermal calorie testing and
Cogan's syndrome, see Chap. 9, p. 154).
pendular body rotation in the dark
Very little work has been done on the long-term
-Increased postural sway with eyes closed and/or standing on foam
prognosis for the different forms of BVF. Recovery of
rubber
hearing and vestibular function is possible in postIncidence/age/sex
meningitic cases, if failure was caused by serous
- Rare condition that manifests throughout life (mean age = 52
labyrinthitis rather than suppurative destructive
years)
labyrinthitis (Fortnum 1982; Rinne et al. 1995;
- Without preference of sex
Bronstein et al. 1995). Partial recovery has also been
Pathomechanism
described in more than 50% of patients with simulProgressive loss of bilaterallabyrinthine and/or vestibular nerve
taneous or sequential idiopathic BVF (Table 8.4;
function due to various aetiologies with concurrent somatosensory
Vibert et al. 1995).
and visual "compensation" (substitution) of vestibular function for
spatial orientation, ocular stabilisation, and postural control
Although increasingly popular (Krebs et al. 1993;
Foster 1994; Herdman 1994), vestibular rehabilitaAetiologies
Ototoxicity, cerebellar degeneration, meningitis, tumours, immunetion has significantly improved functional dynamic
mediated inner ear disease, neuropathies, bilateral sequential
stability during locomotion only in some patients
vestibular neuritis, bilateral Meniere's disease, congenital
with complete bilateral vestibular loss (Krebs et al.
malformation, familial vestibulopathy, vascular disorders, and others
1993; Foster 1994; Telian et al. 1991). Despite these
Idiopathic
(>20%)
objectively disappointing test results, subjective
evaluation of the efficacy of vestibular rehabilitation Course/prognosis
is much more positive (Telian et al. 1991). Most of
BVF may develop simultaneously or sequentially, take an abrupt or
the patients who received physical therapy believed
slowly progressive course, be complete or incomplete. Permanent loss
of vestibular function is most frequent, but partial recovery is possible,
"that the session provided them with new informaparticularly in the idiopathic, post-meningitic, and ototoxic groups
tion that was helpful in understanding and making
adjustments for their condition, even though most of Management
(a) Prevention of BVF (ototoxic drugs)
them had received previous counseHing from a
(b) Recovery from BVF (immune-mediated inner ear disease)
physician" (Telian et al. 1991). Thus, subjective
(c) Vestibular rehabilitation
improvement was much more dramatic than the
objectively achieved improvement in ambulation or Differential diagnosis
(a) Of the various disorders causing BVF
equilibrium.
Table 8.4. ENG controls with a follow-up from 1 to 7 years of patients
with BVF
a. Simultaneous BVF group (n = 8/11)

Persistent bilateralloss (pendular and caloric testing)


Partial recovery
(symmetrie response to pendular testing)
with unilateral response in caloric testing
with partial bilateral response in caloric testing
without response in caloric testing
Complete recovery

4
1
2
1
1

b. Sequential BVF group (n =3/11)

Persistent bilateralloss (pendular and calorie testing)


Partial recovery
in caloric examination with symmetrie pendular testing
in caloric and pendular examinations
Complete recovery
From Vibert et al. (1995).

o
0
3

1
2
0

(b)Of disorders similar in symptomatology (unsteadiness and


oscillopsia)
- Cerebellar or ocular motor disorders without BVF
- Phobie postural vertigo
-Intoxication
- Vestibular paroxysmia
- Perilymph fistula
- "Cervieal vertigo"
- Orthostatic hypotension
- Hyperventilation syndrome
- Visual disorders
- Unilateral vestibular lass

Bilateral vestibulopathy

References
Allum IHI, Honegger F, Schicks H (1994) The influence of a bilateral peripheral vestibular deficit on postural synergies. I Vestib
Res 4:1-22
Agusti C, Ferran F, Gea I, Picado C (1991) Ototoxic re action to erythromycin. Arch Intern Med 151:380
Arbusow V, Strupp M, Dieterich M, Stcker W, Naumann A, Schulz
P, Brandt Th (1998) Serum antibodies against membranous
labyrinth in patients with "idiopathic" bilateral vestibulopathy. I
NeuroI245:l32-l36
Baloh RW (1984) Dizziness, hearing loss, and tinnitus: the essentials of neurootology. Davis, Philadelphia
Baloh RW, Honrubia V, Yee SD, Hess K (1984) Changes in the
human vestibular-ocular reflex after loss of peripheral sensitivity. Ann NeuroI16:222-228
Baloh RW, Iacobson K, Honrubia V (1989) Idiopathic bilateral
vestibulopathy. Neurology 39:272-275
Baloh RW, Iacobson K, Fife T (1994) Familial vestibulopathy: a
new dominantly inherited syndrome. Neurology 44:20-25
Barna B, Hughes G (1988) Autoimmunity and otologic
disease:clinical and experimental aspects. Clinics Labor Med
8:385-398
Barza M, Lauermann M (1978) Why monitor serum levels of gentamicin? Cl in Pharmacokin 3:202-215
Bhansali SA, Stockwell CW, Bojrab Dl (1993) Oscillopsia in
patients with loss of vestibular function. Otolaryngol Head
Neck Surg 109:120-125
Bicknell IM, Holland IV (1978) Neurologic manifestations of
Cogan syndrome. Neurology (Minneap) 28:278-281
Bles W, Klren Th, Bchele W, Brandt Th (1983) Somatosensory
nystagmus: Physiological and clinical aspects. Adv Oto- RhinoLaryngoI30:30-33
Bhmer A, Fisch U (1993) Bilateral vestibular neurectomy for
treatment ofvertigo. Otolaryngol Head Neck Surg 109:101-107
Brandt Th (1996) Bilateral vestibulopathy revisited. Eur I Med Res
1:361-368
Brandt Th, Dieterich M (1988) Oscillopsia and motion perception.
In: Kennard C, Clifford-Rose F (eds) Physiological aspects of
clinical neuro-ophthalmology. Chapman & Hall, London, pp
321-339
Brandt Th, Dieterich M (1996) Postural imbalance in peripheral
and central vestibular disorders. In: Bronstein AM, Brandt Th,
and Woollacott M (eds) Clinical disorders of balance posture
and gait. Edward Arnold, London, pp l31-146
Brandt Th, Strupp M (1992) Otoneurology. Curr Opin Neurol
Neurosurg 5:727-732
Bronstein AM, Hood DI (1986) The cervico-ocular reflex in normal subjects and patients with absent vestibular function. Brain
Res 373:399-408
Bronstein AM, Gresty AM, Mossman SS (1992) Pendular pseudonystagmus arising as a combination of head tremor and
vestibular failure. Neurology 42: 1527 -1531
Bronstein AM, Morland AB, Ruddock KH, Gresty MA (1995)
Recovery from bilateral vestibular failure: implications for visual
and cervico-ocular function. Acta Otolaryngol (Stockh) Suppl
520:405-407
Brookes GB, Gresty MA, Nakamura T, Metcalfe T (1993) Sensing
and controlling rotational orientation in normal subjects and
patients with loss of labyrinthine function. Am I Otol
14:349-351
Bttner U, Ott M, Heimchen Ch, Yousry T (1995) Bilateralloss of
eighth nerve function as the only clinical sign of vertebrobasilar dolichoectasia. I Vestib Res 5:47-51
Chambers BR, Mai M, Barber HO (1985) Bilateral vestibular loss,
oscillopsia, and the cervico-ocular reflex. Otolaryngol Head
Neck Surg 93:403-407

139
Cheung BK, Howard IP, Nedzelski IM, Landolt IP (1989)
Circularvection about earth-horizontal axes in bilateral
labyrinthine-defective subjects. Acta Otolaryngol (Stockh)
108:336-344
Cody DTR, Williams HL (1960) Cogan's syndrome. Laryngoscope
70:447-478
Cogan DG (1945) Syndrome of nonsyphilitic interstitial keratitis
and vestibuloauditory symptoms. Arch Ophthalmol 33: 144-149
Cote DN, Molony TB, Waxman I, Parsa D (1993) Cogan's syndrome manifesting as sudden bilateral deafness: diagnosis and
management. South Med I 86:1056-1060
Dandy WE (1941) The surgical treatment of Meniere's disease.
Surg Gynecol Obstet 72:421-425
Diamant H (1946) Sound localisation and its determination in
connection with some cases of severely impaired function of
vestibular labyrinth, but with normal hearing. Acta Otolaryngol
(Stockh) 34:576-586
Dichgans I, Brandt Th (1978) Visual-vestibular interaction: effects
on self-motion perception and postural contro!. In: Held R,
Leibowitz HW, Teuber H-L (eds) Handbook of sensory physiology, vol 8. Perception. Springer, Berlin Heidelberg New York, pp
755-804
Dichgans I, Held R, Young LR, Brandt Th (1972) Moving visual
scenes influence the apparent direction of gravity. Science
178:1217-1219
Dichgans I, Bizzi E, Morasso P, Tagliasco V (1973) Mechanisms
underlying recovery of eye-head co ordination following
labyrinthectomy in monkeys. Exp Brain Res 18:548-562
Enbom H, Magnussson M, Pyykk I (1991) Postural compensation in children with congenital or early acquired bilateral
vestibular loss. Ann Otol Rhinol Laryngoll00:472-478
Feneley MR, Murthy P (1994) Acute bilateral vestibulo-cochlear
dysfunction following occipital fracture. I Laryngol Otol
108:54-56
Fortnum HM (1982) Hearing impairment after bacterial meningitis.Arch Dis Child 67:1128-1l33
Foster CA (1994) Vestibular rehabilitation. Bailliere's Clin Neurol
3:577-592
Frohman EM, Tusa R, Mark AS, Cornblath DR (1996) Vestibular
dysfunction in chronic inflammatory demyelinating polyneuropathy. Ann NeuroI39:529-535
Gill GV, Bell DR (1982) Persisting nutrition al neuropathy amongst
former war prisoners. I Neurol Neurosurg Psychiatry
45:861-865
Glasauer S, Amorim MA, Vitte E, Berthoz A (1994) Goal-directed
linear locomotion in normal and labyrinthine-defective subjects. Exp Brain Res 98:323-335
Grad A, Baloh RW (1989) Vertigo ofvascular origin. Clinical and
electronystagmographic features in 84 cases. Arch Neurol
46:281-284
Graybiel A, Smith CR, Guedry FE, Miller EF, Fregly AR, Cramer DB
(1972) Idiopathic progressive vestibular degeneration. Ann Otol
Rhinol LaryngoI81:165-179
Grimaldi LME, Luzi L, Martino GV, Furlan R, Nemni R, Antonelli
A, Canal N, Pozza G (1993) Bilateral eighth cranial nerve neuropathy in human immunodeficiency virus infection. I Neurol
240:363-366
Grnbauer WM, Dieterich M, Brandt T (1998) Bilateral vestibular
failure impairs visual motion perception even with the head
still. NeuroReport 9:1807-1810
Guttich H, Stark R (1965) Doppelseitiger, praktisch vollkommener,
wahrscheinlich angeborener Vestibularisausfal!. HNO
l3:177-180
Halmagyi GM, Curthoys IS (1988) A clinical sign of canal paresis.
Arch NeuroI45:737-739
Hanson W, Parnes LS (1995) Vestibular nerve compression in
Camurati-Engelmann disease. Ann Otol Rhinol Laryngol
104:823-825

140
Harris EN, Asherson RA, Hughes GRV (1988) Antiphospholipid
antibodies: autoantibodies with a difference. Ann Rev Med
39:261-271
Husler R, Vibert D, Koralink I], Hirschul BC (1991) Neurootological manifestation in different stages of HIV infection.
Acta Otolaryngol (Stockh) 481:515-521
Haynes BF, Kaiser-Kupfer MI, Mason I, Fanci AS (1980) Cogan
syndrome: studies in 13 patients, long-term follow-up and a
review of the literature. Medicine 59:426-441
Heimbrand S, Bronstein AM, Gresty MA, Faldon ME (1996)
Optically induced plasticity of the cervico-ocular reflex in
patients with bilateral absence of vestibular function. Exp Brain
Res 112:372-380
Herdman SJ (1994) Vestibular rehabilitation. Davis, Philadelphia
Herdman SJ, Sandusky AL, Hain TC, Zee DS, Tusa RJ (1994)
Characteristics of postural stability in patients with aminoglycoside toxicity. I Vestib Res 4:71-80
Horak FB, Shupert CL, Dietz V, Horstmann G (1994) Vestibular
and somatosensory contributions to responses to head and
body dis placements in stance. Exp Brain Res 100:93-106
Hu DH, Qiu WQ, Wu BT, Fang LZ, Zhou F, Gu YP, Zhang QH, Yan
IH, Ding YQ, Wong H (1991) Genetic aspects of antibiotic
induced deafness: mitochondrial inheritance. I Med Genet
28:79-83
Hughes GB, Kinney SE, Barna BP, Calabrese LH (1984) Practical
versus theoretical management of autoimmune inner ear disease. Laryngoscope 94:758-767
Huygen PLM, Verhagen WIM (1994) Peripheral vestibular and
vestibulo-cochlear dysfunction in hereditary dis orders. A
review of the literature and areport on some additional findings. I Vestib Res 4:81-104
Ingles IT, Macpherson IM (1995) Bilaterallabyrinthectomy in the
cat: effects on the postural response to translation. I
NeurophysioI73:1181-1191
lger L, Strupp M, Brandt T, Reiser M (1997) Bildgebung von
Labyrinth und Nervus vestibularis. Nervenarzt 68:443-458
lahrsdoerfer RA, Thompson EG, Johns MM, Cantrell RW (1981)
Sarcoidosis and fluctuating hearing loss. Ann Otol Rhinol
LaryngoI90:161-163
Johnson L, Hawkins I (1972) Sensory and neural degeneration
with ageing, as seen in microdissections of the human inner
ear. Ann Otol Rhinol Laryngol81:179
Kasai T, Zee DS (1978) Eye-head co ordination in labyrinthine
defective human beings. Brain Res 144:123-141
Kommune S, Nogami K, Inoue H, Uemura T (1993) Bilateral
Mondini dysplasia with normal hearing. ORL I
Otorhinolaryngol Relat Spec 55: 143-146
Knigsmark BW, Gorlin RI (1976) Genetic and metabolic
deafness. Saunders, Philadelphia
Krebs DE, Gill-Body KM, Riley PO, Parker SW (1993) Doubleblind, placebo-controlled trial of rehabilitation for bilateral
vestibular hypofunction: preliminary report. Otolaryngol Head
Neck Surg 109:735-741
Lamont MH, Edwards PT, Windsor RS (1980) Streptococcal
meningitis in pigs: results of a five year survey. Vet Rec
107:467-469
Lenard HG, Voit T, Lamprecht A, Kahn T, Neuen-Iacob E,
Ruitenbeek W (1992) Sudden loss of hearing and vestibular
function, muscular weakness, and multiple white matter lesions
in preschool children. Neuropediatrics 23:221-224
Lu CB, Schuknecht HF (1994) Pathology of prelingual profound
deafness: magnitude of labyrinthitis fibro-ossificans. Am I Otol
15:74-85
Magnusson M, Pyykk I (1991) Postural compensation in children
with congenital or early acquired bilateral vestibular loss. Ann
Otol Rhinol LaryngoI100:472-478
Magnusson M, Padoan S, Karlberg M, lohansson R (1991) Delayed
onset of ototoxic effects of gentamicin in treatment of

Vertigo
Meniere's disease. Acta Otolaryngol (Stockh) Suppl
481:610-612
Mazzoni A (1974) Internal auditory artery supply to the petrous
bone.Ann Otol Rhinol LaryngoI81:13-21
McDonald Tl, Vollertsen RS, Younge BR (1985) Cogan's syndrome:
Audiovestibular involvement and prognosis in 18 patients.
Laryngoscope 95:650-654
McGarth JH, Barber HO, Stoyanoff S (1989) Bilateral vestibular
loss and oscillopsia. I OtolaryngoI18:218-221
Merchant SN, Schuknecht HF (1988) Vestibular atelectasis. Ann
Otol Rhinol LaryngoI97:565-576
Mller C, dkvist LM (1989) The plasticity of compensatory eye
movements in bilateral vestibular loss. Acta Otolaryngol
(Stockh) 108:345-354
Moscicki RA (1994) Immune-mediated inner ear disorders.
Bailliere's Clin NeuroI3:547-563
Myers E, Bernstein I, Fostiropolous G (1965) Salicylate ototoxicity.
A clinical study. N Engl J Med 273:587
Nakamura T, Bronstein AM (1995) The perception of head and
neck angular displacement in normal and labyrinthinedefective subjects: a quantitative study using a "remembered
saccade" technique. Brain 118:1157-1168
Nance WE, Sweeney A (1975) Genetic factors in deafness in early
life. Otolaryngol Clin North Am 8:19
Nuti D, Pas sero S, DiGirolamo S (1996) Bilateral vestibular loss in
vertebrobasilar dolichoectasia. I Vestib Res 6:85-91
Ogata Y, Sekitani T, Shimogori H, Ikeda T (1993) Bilateral vestibular neuronitis.Acta Otolaryngol (Stockh) SuppI503:57-60
Paige GD (1992) Senescence of human visual-vestibular interactions. 1. Vestibulo-ocular reflex and adaptive plasticity with
ageing. J Vestib Res 2:133-151
Paulus W, Straube A, Brandt Th (1987) Visual postural performance after loss of somatosensory and vestibular function. I
Neurol Neurosurg Psychiatry 50:1542-1545
Pinsky L, DiGeorge AM (1966) Congenital familial sensory neuropathy with anhidrosis. J Pediatr 68: 1-l3
Rinne T, Bronstein AM, Rudge P, Gresty MA, Luxon LM (1995)
Bilateralloss of vestibular function. Acta Otolaryngol (Stockh)
SuppI520:247-250
Rowe-Jones IM, Macallan DC, Sorooshian M (1990) Polyarteritis
nodosa presenting as bilateral sudden onset cochleo-vestibular
failure in a young woman. I Laryngol Otoll04:562-564
Sato K (1982) Histopathological study on the vestibular toxicity of
six aminoglycoside antibiotics. Drugs Exp Clin Res 8:259
Schuknecht HF (1974) Pathology of the ear. Harvard University
Press, Cambridge, Mass
Schuknecht HF (1980) Mondini dysplasia. A clinical and pathological study. Ann Otol Rhinol LaryngoI85:1-23
Schuknecht HF, Witt RL (1985) Acute bilateral sequential vestibular neuritis. Am I OtolaryngoI6:255-257
Simmons FB (1993) Patients with bilateral loss of caloric
response.Ann OtoI82:175-178
Suzuki M, Kitahara M (1992) Immunologie abnormality in
Menii~re's disease. Otolaryngol Head Neck Surg 107:57-62
Telian SA, Shepard NT, Smith-Wheelock M, Hoberg M (1991)
Bilateral vestibular paresis: diagnosis and treatment.
Otolaryngol Head Neck Surg 104:67-71
Terjung B, Heimchen C, Samtleben W (1993) Glucocorticoid
monotherapy for Cogan's syndrome? Dtsch Med Wochenschr
118:1231-1235
Thomas JE, Cody DTR (1981) Neurologic perspectives of otosclerosis. Mayo Clin Proc 56: 17
Thomson DB, Inglis JT, Schor RH, Macpherson JM (1991) Bilateral
labyrinthectomy in the cat: motor behaviour and quiet stance
parameters. Exp Brain Res 85:364-372
Trop I, Schloss MD, Polomeno R, Der-Kaloustian V (1995) Usher
syndrome in four siblings from a consanguineous family of
Pakistani origin. J Otolaryngol 24: 102-1 04

Bilateral vestibulopathy
Tsunoda I, Kanno H, Watanabe M, Shimoji S, Hirayama K, Sumita
H, Yamamoto T (1994) Acute simultaneous bilateral vestibulocochlear impairment in neuro-Beh.;et's disease. Auris Nasus
Larynx 21 :243-24 7
Verhagen WIM, Huygen PLM, Horstink NWIM (1987) Familial
congenital vestibular areflexia. J Neurol Neurosurg Psychiatry
50:933-935
Vibert 0, Liard P, Husler R (1995) Bilateral idiopathic loss of
peripheral vestibular function with normal hearing. Acta
Otolaryngol (Stockh) 115:611-615
Vollertsen RS, McDonald TJ, Younge BR, Banks PM, Stanson AW,
Ilstrup DM (1986) Cogan's syndrome: 18 cases and a review of
the literature. Mayo Clin Proc 61:344-361
Waespe W, Schwarz U, Wolfenberger M (1992) Firing characteristics of vestibular nuclei neurons in the alert monkey after bilateral vestibular neurectomy. Exp Brain Res 89:311-322

141
Waterston JA, Barnes GR, Grealy MA, Luxon LM (1992)
Co ordination of eye and head movements during smooth pursuit in patients with vestibular failure. J Neurol Neurosurg
Psychiatry 55:1125-1131
Wester DC,Atkin CL, Gregory MC (1995) Alport syndrome: clinical update. J Am Acad AudioI6:73-79
Yliskoski JS, House JW, Hernadez 1(1981) Eighth nerve alcoholic
neuropathy: A case report with light and electron microscopic
findings. J Laryngol Otol 95:631-642
Young 0, Eldridge R, Gardner W (1970) Bilateral acoustic neuroma in a large kindred. JAMA 214:347
Zee OS (1978) Ophthalmoscopy in examination of patients with
vestibular dis orders. Ann NeuroI3:373-374

Miscellaneous vestibular nerve and


labyrinthine disorders

Unilateral or bilateral vestibular dysfunctions can be


produced by conditions other than the common disorders discussed in the previous chapters. These can
be conveniently cJassified under the following headings: congenital (see also Chap. 26, Hereditary disorders), infectious, autoimmune, vaseular (see Chap.
7, Vestibular paroxysmia and Seetion E, Vascular vertigo), and toxie causes (see also Chap. 24, Iatrogenie
vestibular dis orders and Chap. 28, Drugs and vertigo),
as weil as tumours (see also Chap. 15), and trauma
(see also Chaps. 22 and 23). This ehapter gives
examples of various eonditions that manifest with
vertigo and disequilibrium and are due to the
involvement of the peripheral vestibular system.
Imaging of the inner ear is important for the differential diagnosis of these eonditions.

Imaging of the labyrinth and


vestibular nerve
High-resolution magnetie resonanee imaging (HRMRI) and eomputed tomography (HR-CT) of the
inner ear are becoming more important for the diagnosis of peripheral vestibular lesions (Seltzer and
Mark 1991; Mark et al. 1992; Mafee 1993; Casselman
1994; Casselman et al. 1994; Jger et al. 1997).
Modern HR-MRI techniques allow the visualisation
of detailed anatomical features of the vestibulocochlear regions (Figs 9.1-9.3) as weIl as pathological changes caused by disorders of the inner ear such
as neoplastic formations (e.g. small intracanalicular
acoustic neuromas), anomalies causing vertigo
and hearing loss (e.g. Mondini's malformation,
perilymph fistula, vestibular paroxysmia), and
inflammatory diseases (e.g. Cogan's syndrome,
labyrinthitis, zoster neuritis). HR-CT is still the first
examination that should be performed in patients
with middle ear diseases (e.g. tumour, infection),
trauma (e.g. temporal bone fractures), or fibro-

osseous diseases. While the quality of imaging of the


vestibulocochlear system has dramatically improved
in recent years, there are still several peripheral
vestibular dis orders that cannot yet be visualised,
e.g. benign paroxysmal positioning vertigo, idiopathic vestibular neuritis, or Meniere's disease (Jger
et al. 1997).

Congenital causes
In the majority of cases the diagnosis of hereditary
deafness with and without partial or complete
vestibular loss rests on a positive family history (see
Chap. 26). Alport's, Usher's, and Waardenburg's syndromes (p. 378) usually cause bilaterallabyrinthine
deficiency if they affect the vestibular system (Nance
and Sweeney 1975; Knigsmark and Gorlin 1976;
Huygen and Verhagen 1994; Verhagen and Huygen
1994). Congenital vestibular loss is secondary to
either abnormal genetic or intrauterine factors
including infection, intoxication, or anoxia. Among
the children of mothers who suffer from rubella during the first trimester of pregnancy, 50-70% will
have hearing loss and a smaller proportion will also
have vestibular impairment (Barr and Lundstrm
1961). The Scheibe syndrome, a cochleosaccular
malformation with sparing of the semicircular
canals and the utricle, can also result from a rubella
infection. The most frequently occurring congenital
viral infections in utero are those produced by
cytomegalovirus. Subclinical cytomegalovirus is the
most common viral agent causing sensorineural
hearing loss (with and without vestibular impairment) among pediatric patients (Pappas 1983).
Thalidomide taken during pregnancy can cause
aplasia of the inner ear as well as other ear and body
deformities such as dysmelia. Various genetic and
acquired factors playa role in Mondini dysplasia
(Fig. 9.4), which involves malformation of the

143

Vertigo

144

145

Miscellaneous vestibular nerve and labyrinthine disorders

Fig.9.2. Imaging the ductus and saccus endolymphaticus by high resolution MRI. a Axial projection, T2 -weighted 3D-CISS. The
endolymphatic duct (arrow) can be delineated as a structure with high signal extending from within the vestibule to the endolymphatic sac (curved arrow). b Sagittal projection, T2 -weighted 3D-ClSS. The endolymphatic duct (short arrow) can be imaged from the
common crus (long arrow) to the endolymphatic sac. (From Jger et al. 1997.)

membranous and osseous labyrinth with combined


auditory and vestibular loss and may be partial or
complete (Illum 1972; Schuknecht 1980; Komune et
al. 1993). Perilymph fistulas (p. 99) are also considered typical complications, because they frequently
result from stapes deformities. Osteopetrosis
(Albers-Schnberg disease) may compress the
vestibulocochlear nerve (Fig. 9.5). The Klippel-Feil
syndrome, an abnormal fusion of cervical vertebrae,
can be associated with deafness and some vestibular
loss due to temporal bone malformation (McLay and
Maran 1969). Finally, sickle cell anaemia, a chronic
haemolytic anaemia occurring primarily in blacks

and characterised by sickle-shaped red blood cells


due to homozygous inheritance of HbS, can cause
peripheral audio-vestibular dysfunction (Savundra
1996; Savundra et al. 1996).

Infectious causes
These cases comprise viral, bacterial, and specific
(syphilitic, tuberculous) processes involving either
the labyrinth or the vestibular nerve (Table 9.1) .

...
Fig.9.1. Imaging labyrinth and vestibular nerve by high resolution MRI. a Axial projection, T1-weighted 2D-FLASH (fast low angle
shot). The facial nerve with its intrameatal, labyrinthine, and tympanal portion (Iong orrows) is shown. The geniculate ganglion (short
arrow). the greater superficial petrosal nerve (small arrow). and the superior part of the vestibular nerve (curved arrow) can be delineated. The vestibule (open arrow) can also be seen. b Axial projection, T1-weighted 2D-FLASH (fast low angle shot). The vestibulocochlear nerve with its cochlear nerve (large arrow) and the inferior part of the vestibular nerve are imaged. The tympanal portion of
the facial nerve is also demonstrated (short arrows) . c Axial projection T2-weighted 3D-CISS (constructive interference in steady state).
The facial nerve (black arrow) and the superior part of the vestibular nerve (white arrow) can be delineated in the internal auditory
canal. They are surrounded by the hyperintense cerebrospinal fluid . d Axial projection T2-weighted 3D-CISS. The vestibulocochlear
nerve with its cochlear nerve (Iarge white arrow) and the inferior part of the vestibular nerve (small black arrows) are imaged with low
intensity. The intralabyrinthine spaces filled with endo- and perilymph are imaged with high intensity. The vestibule (curved arrow) and
the posterior semicircular canal (open arrow) can also be delineated. The osseous spiral lamina (arrowhead) separates the scala vestibuli (sm all short white arrow) from the scala tympani (smalliong white arrow). The modiolus (short black arrow) is shown as a course,
without signal surrounded by the hyperintense Iymphatic space of the different turns of the cochlea. (From Jger et al. 1997.)

146

Vertigo

c
Fig.9.3. Three-dimen sional imaging of the labyrinth. a 3D-MIP (maximum intensity projection) reconstruction of the 3D-ClSS
images. The cochlea with the osseous spiral lamina (small black arrow), the scala tympani (smalliong white arrow) and the scala
vestibuli (small short white arrow), the vestibule (curved black arrow), the lateral semicircular (large long white arrow), and the posterior
semicircular canal are shown. b 3D-MIP reconstruction of 3D-CISS images. The 2~ turns of the cochlea can be delineated with the
osseous spiral lamina (black arrow), the scala tympani (long white arrowl, and the scala vestibuli (short white arrow). c 3D-MIP reconstruction of 3D-ClSS images.The semicircular canals with the common crus are shown. (From Jger et al. 1997.)

Viral vestibular neuritis (p. 67), sudden deafness,


and Bell's palsy are the most eommon diseases. They
may reeur (Ishiyama et al. 1996) and eombine to
some extent in their manifestation (Watanabe et al.
1993). Sehukneeht (1985) has deseribed a variety of
viral infeetions of the peripheral auditory and
vestibular system (Table 9.2), whieh may be followed
by delayed endolymphatie hydrops (p. 88) that with
time eauses symptomatie Meniere's syndrome. HIV
infeetion ean affeet the vestibular and auditory systems at different stages of the disease (Husler et al.

1991) and lead to eighth nerve neuropathy (Grimaldi


et al. 1993). The spumaretrovirus may be neuropathie to the eighth nerve (Pyykk et al. 1997) as are the
related retroviruses, the human immunodeficieney
and the human T-eellieukaemia viruses.

Herpes zoster oticus


This infeetion (also known as Ramsay Hunt syndrome) manifests as a burning pain in the ear and

147

Miscellaneous vestibular nerve and labyrinthine disorders

Fig.9.4. Mondini syndrome with dysplastic cochlea. a Axial projection, T2-weighted 3D-ClSS. b,c Axial projection, T2-weighted 3D(ISS. Lateral semicircular canal is missing or hypoplastic on both sides. (From Jger et al. 1997.)
Table 9.1 . Frequency and mechanism of vertigo with infections of the
ear and temporal bone
Infection

Vertigo'

Otitis externa

Never

Malignant externaiotitis

Rare

Acute otitis media

Occasionally

Mastoiditis

Occasionally

Cholesteatoma

Occasionally

Petrositis

Occasionally

Viral labyrinthitis

Usually

Bacteriallabyrinthitis

Always

Otosyphilis

Frequently

Herpes zoster oticus

Occasionally

Usual mechanism

Extension of infection to
labyrinth, internal auditory
canal, or both
Toxins enter through round
window
Extension of infection to
the inner ear
Erosion of bony horizontal
semicircular canal
Infection of eighth cranial
nerve
Inflammation of the
membranous labyrinth
Necrosis of the
membranous labyrinth
Osteitis with destruction of
the otic capsule;
endolymphatic hydrops
Inflammation of the eighth
cranial nerve

' Rare, <5%; occasionally, - 20%; frequently, - 50%; usually, > 70%.
From Canalis (1996).

Table 9.2. Neurolabyrinthitis: classification of viral disorders of the


peripheral auditory and vestibular system and sequela of delayed
endolymphatic hydrops

Acute viral labyrinthitis


Acute cochlear labyrinthitis (sudden deafness)
Acute vestibular labyrinthitis (cannot be differentiated clinically from
vestibular neuritis)
Acute cochleovestibular labyrinthitis (rare, simultaneous hearing 1055 and
vertigo)
Acute viral neuritis
Acute cochlear neuritis (rare?)
Acute vestibular neuritis (frequent)
Acute cochleovestibular neuritis
(e.g. herpes zoster)
Delayed endolymphatic hydrops (secondary Meniere's syndrome)
Ipsilateral (to initial hearing 1055)
Contra lateral (to initial hearing 1055)
Bilateral
Modified from Schuknecht (1985).

148

Vertigo

b
Fig.9.5. Osteopetrosis (Albers-5chnberg disease). a Axial projection, T2 -weighted 3D-(155. Narrowing of the internal auditory canal
(arrow) can be detected as a recess of the perineuralliquor space. b Axial projection, T,-weighted 2D-FLA5H. As a result of narrowing of
the internal auditory canal, the vestibulocochlear and facial nerve cannot be differentiated (arrow). (From Jger et al. 1997.)

the mastoid region, followed by a vesicular eruption


within or near the external auditory canal.
Combined auditory and vestibular loss and, in addition, facial nerve palsies may develop. Infiammation
of the geniculate ganglion is not the cause, as Hunt
originally supposed (1908), but rather viral neuritis
of the seventh and eighth cranial nerves (Denny
Brown et al. 1944; Blackley et al. 1967; Zajtchuk et al.
1972). Infectious involvement of the vestibular nerve
may be partial, sparing posterior semicircular canal

and part of the saccule (Proctor et al. 1979). For a


more detailed discussion, see vestibular neuritis
(p. 67). MRI scans may reveal contrast enhancement
in either or both (Fig. 9.6) parts (pars superior and
inferior) of the vestibular nerve due to viral infiammation. Herpes zoster neuritis involving superior
and inferior parts of the vestibular nerve causes ocular tilt reaction (Arbusow et al. 1998a). Herpes zaster
oticus may also lead to labyrinthitis (Fig. 9.7) .
Benign paroxysmal positioning vertigo, a typical

149

Miscellaneous vestibular nerve and labyrinthine disorders

Fig.9.6. Herpes zaster oticus (Ramsay Hunt syndrome). Coronal projection, T,-weighted, 2D-FLASH sequence reveals gadolinium
enhancement of pars superior and inferior of right vestibular nerve (arrows). (From Arbusow et al. 1998a.)

Fig.9.7. Herpes zoster oticus with labyrinthitis. Coronal projection, T,-weighted 2D-FLASH, post Gd-DTPA. Enhancement in vestibule
(/arge arrow), lateral (5mallshort arrow), and anterior (smalliong arrow) semicircular canals can be detected. (From Jger et al. 1997.)

sequel of vestibular neuritis (p. 67), has also been


mentioned in connection with Ramsay Hunt syndrome (Longridge 1989), but there is no eonvincing
evidenee of a eausal inter relation. In mumps (epidemie parotitis), deafness (usually unilateral and
reversible) oeeurs in about 4% of adult patients.
Coneomitant vestibular damage is evident on ealorie
testing, but is easily overlooked (Hyden et al. 1979).

Acute otitis media


Aeute otitis media (suppurative or serous) frequently
eauses eonduetive hearing loss in ehildren (Goodhill
1979). MRI seans ean visualise otitis media with
assoeiated toxie labyrinthitis (Fig. 9.8). Invasion of
the labyrinth from the middle ear is rare, but if it
oeeurs, it usually eauses spontaneous tympanie

150

Vertigo

c
Fig.9.8. Otitis media with associated toxic labyrinthitis.a Axial projection, T,-weighted 2D-FLASH.As a sign of subacute haemorrhage hyperintense lesions are seen in cochlea (smalllong arrow), vestibule (smallshort arrow) and in posterior semicircular canal
(large short arrow). b Axial projection, T,-weighted 2D-FLASH, post Gd-DTPA. Enhancement can be delineated in cochlea (long arrow)
and in vestibule (short arrow).c Axial projection, T,-weighted 2D-FLASH, post Gd-DTPA. Enhancement can be delineated in cochlea
(smallshort arrow), in vestibule (large long arrow), and in cochlear nerve (smalllong arrow). (From Jger et al. 1997.)

Miscellaneous vestibular nerve and labyrinthine disorders

membrane rupture with otorrhea. Bacterial labyrinthitis may develop from bacterial meningitis via
the perilymphatic space (Fig. 9.9). It subsequently
destroys the labyrinth, resulting in irreversible
vestibular and auditory loss (Schuknecht 1974).
Histopathological examination of children who died
of meningitis revealed that concurrent acute otitis
media did not cause the bacterial meningitis; it
appeared to be the result of retrograde bacterial
invasion from the meninges (Eavey et al. 1985).

Specific infections
Tuberculous labyrinthitis is also more often a complication of tuberculous meningitis than of tuberculous otitis media. Syphilitic labyrinthitis may be
either congenital (more common, with typical stigmata of congenital syphilis) or acquired. The peak
incidence for the former is around the fourth to fifth
decades, and the latter, around the fifth and sixth
decades. It typically progresses slowly with some
fiuctuating episodes of hearing loss and vertigo. The
spontaneous course and the histologie al findings of
endolymphatic hydrops with atrophy and loss of
neurons are similar to those in Meniere's disease
(Schuknecht 1974; Kobayashi et al. 1991). Borrelia
infections have also been reported to cause
vertigo/nystagmus (Rosenhall et al. 1988); they may
mimic vestibular neuritis and cause sudden deafness
(Ishizaki et al. 1993; Hyden et al. 1995). Perilymph
fistulas may result from bacterial, syphilitic, or
tuberculous infections (p. 102).

151

Cholesteatoma
Chronic otomastoiditis can cause a cholesteatoma to
develop in the temporal bone. A cholesteatoma
invades the middle ear by perforating the tympanic
membrane and may even affect the horizontal semicircular canal with secondary fistualisation. This
cyst -like structure contains dis integration products,
predominantly keratin, and is characterised by
chronic expansion, epithelial ingrowth, and bone
destruction. The expansion of the cholesteatoma,
secretion of bacterial endotoxins, and induction of
osteoclastic activity combine to destroy the bone
(Canalis 1996). Cholesteatomas can be visualised
with CT (Mafee 1993) and MRI scans (Figs.
9.10-9.12). The sole treatment is surgery.

Autoimmune inner ear disorders


Immune-mediated sensorineural hearing and
vestibular loss comprise a frequently undetected,
heterogeneous collection of autoimmune disorders,
which are clinically relevant because of their potential treatability. Immune-mediated inner ear disease
is characterised by sensorineural hearing loss that is
most often rapidly progressive and asymmetrieal,
and for which no cause can otherwise be identified
(Moscicki 1994). Bilateral sensorineural hearing loss
may be accompanied by vestibular symptoms.
Lehnhardt (1958) was the first to suspect that bilateral

Fig.9.9. Bacterial meningitis with accompanying labyrinthitis. Axial projection, T,-weighted 2D-FLASH, post Gd-DTPA. Enhancement
can be delineated in cochlea Uarge long arrowl. in vestibule (curved arrow), lateral semicircular canal (sm all short arrow, posterior semicircular canal (arrowhead), and in meninges in interna I auditory (anal (smalliong arrow). (From Jger et al. 1997.)

152

Vertigo

d
Fig. 9.1 O.

Cholesteatoma at the apex of the pyramid. a Axial projection, T2 -weighted TSE (turbo spin echo). The cholesteatoma has a
high signal. b Axial projection, T,-weighted 2D-FLASH. The cholesteatoma at the apex of the pyramid (arrow) has a slightly higher signal than the labyrinth. c,d Axial and coronal projections, T,-weighted 2D-FLASH, post Gd-DTPA. The cholesteatoma at apex of the pyramid has a surrounding rim of enhancement (arrow). (From Jger et al. 1997.)

sensorineural hearing loss may have an immunemediated pathology. Later McCabe (1979, 1989) was
able to demonstrate a positive response to immunosuppression and gave indirect evidence supporting
such a diagnosis. Far other arguments attesting to
the existence of auto immune inner ear disarders, see
Table 9.3. Moreover, experimental evidence even
indicates that the inner ear can function as an
immune organ (Far review see Moscicki 1994; Harris
and O'Driscoll1996). An animal model of these disorders has been established in the guinea-pig
(Harris 1987), but only limited aspects can be
extrapolated to humans. Serum antibodies to inner
ear tissue have been repeatedly detected (Arnold et
al. 1985) by different methods, but no clear-cut, clinical correlation has been established (Arnold and
Pfaltz 1987; Hughes et al. 1984). Eight of twelve
patients with "idiopathic" bilateral vestibulopathy
(p. 132) exhibited antibodies against membranous
labyrinth (ampulla, semicircular canal, saccule, and
utricle) (Arbusow et al. 1998b). Although this finding
may be an epiphenomenon, a small subgroup of
argan-specific autoantibodies could synergise with a

cellular response to develop vestibular loss. Western


blot immunoassays have been used to find a laboratory marker far immune-mediated inner ear disease
which can serve as a predictor for responsiveness to
steroids (Moscicki et al. 1994).
Bilateral sympathetic organ dysfunction is a major
concern in ophthalmology (Chan et al. 1995). Injury
to one eye may expose antigens previously concealed
from reactive T-cells and thus trigger an autoimmune process, which also seriously damages the
contralateral, intact eye. Contralateral hearing loss
observed after ear surgery has been attributed to the
same pathomechanism (Harris et al. 1985). Pyykk
et al. (1997) followed 478 patients who underwent
operations for a unilateral vestibular schwannoma.
They found eight in whom significant hearing loss
started 2-4 years after surgery, and in three of these
eight patients, sympathetic cochleolabyrinthitis was
suspected, because the clinical course was identical
to that of delayed hydrops. Schulz et al. (1999)
described the first case of sympathetic contralateral
vestibulopathy following unilateral zoster oticus.
They suggest that an auto immune pathogenesis of

153

Miscellaneous vestibular nerve and labyrinthine disorders

b
Fig.9.11. Cholesteatoma in the middle ear.a Axial projection, T1-weighted 2D-FLASH. The cholesteatoma in the middle ear with
hypointense signal (arrow). b Axial projection, T1-weighted 2D-FLASH, post Gd-DTPA. A surrounding rim of enhancement can be delineated around the cholesteatoma (arrow). (From Jger et al. 1997.)

the contralateral vestibular failure is more likely


than bilateral varicella zaster infection because of
(a) no evidence of vesicular eruptions on the left
auride and the virtual absence of antiviral antibodies
after onset of bilateral vestibulopathy, (b) the prompt
response of the contralateral vestibule to immunosuppressive therapy, and (c) the presence of atypical
nerve tissue-specific autoantibodies against a 45kDa pro tein.
Autoimmune inner ear disease should be suspected
whenever progressive bilateral sensorineural hearing loss and/or vestibular loss remain unexplained
despite extensive diagnostic examinations (Moscicki

1994; Harris and O'Driscoll 1996). Steroids (e.g.


prednisone, 60 mg daily) are the first choice for an
empirical therapy. Indications for more aggressive
treatments with cydophosphamide, azathioprine, or
methotrexate depend on the strength of dinical and
laboratory evidence for an autoimmune disease
and on the course of the dis order. Plasmapheresis
and high-dose intravenous gammaglobulins have
also been tried on an empirical basis but not within
the framework of controlled studies.
Vestibular and sensorineural hearing loss may be
part of the multiple manifestations of a number
of well-defined systemic autoimmune diseases

154

Vertiga

b
Fig.9.12.

Cholesterol cyst of the middle ear. a Axial projection, T2 -weighted TSE. The cholesterol cyst has a hyperintense signal

(arrow). b Axial projection, T,-weighted 2D-FLASH. The cholesterol cyst has a hyperintense signal. (From Jger et al. 1997.)

Table 9.3.
disease

Immune-mediated mechanisms and clinical features of inner ear

Evidence for
mechanisms

Clinical features of
this disease

-Impraved hearing of patients treated with


immunosuppressive regimens
- Detection of specific cellular and antibody-mediated
immune responses in some patients
-Identification of immunological pracesses involving the
inner ear
- Animal models demonstrate that autoimmunity to
inner ear tissue can result in inner ear disease
- Rapidly progressive course over days, weeks, or months
- Usually bilateral, but onset asymmetrical; may be unilateral
- Vestibular symptoms variably present
- Pressure and tinnitus often present
- More common in middle-aged women, but may occur
in both sexes and at any age

Modified fram Moscicki (1994).

(Iable 9.4) including, fr example, Beh<;:et's disease


(Gemiginani et al. 1991; Isunoda et al. 1994), polyarteritis nodosa (Rowe-Jones et al. 1990), neurosarcoidosis (von Brevern et al. 1997) and Cogan's
syndrome.

(ogan's syndrome
In 1945 Cogan described a syndrome of nonsyphilitic interstitial keratitis and audiovestibular
symptoms. Ihis rare auto immune disease of young
adults res ponds to glucocorticoids and cytotoxicimmunosuppressive agents (Vollertsen et al. 1986;

155

Miscellaneous vestibular nerve and labyrinthine disorders


Table 9.4. Immune-mediated systemic diseases associated with sensorineural hearing 1055
Cogan's syndrome
Polychondritis
Systemic lupus erythematosus
Juvenile rheumatoid arthritis
Adult rheumatoid arthritis
Systemic vasculitis
Polyarteritis nodosa
Wegener's granulomatosis
Giant cell arteritis
Beh~et's disease
From Moscicki (1994).

Terjung et al. 1993). Early treatment, however, is crucial. Episodes of sudden vestibulo-auditory dysfunction resemble Meniere's attacks. The hearing loss,
initially unilateral, subsequently involves both ears,
at which time histology shows diffuse degeneration
of neuronal tissue. The key to the diagnosis is the
dose temporal proximity to flare-ups of interstitial
keratitis, which are associated with photophobia,
lacrimation, and eye pain (Cogan 1945; Haynes et al,
1980). Systemic manifestations indude elevated
erythrocyte sedimentation rate, increased white
blood cell count or C-reactive protein, anaemia,
thrombocytosis, and fever. Serious outcomes indude
deafness and, less frequently, vasculitis, aortitis with
aortic insufficiency, blindness, and death (Cody and
Williams 1960; Haynes et al. 1980; Vollertsen et al.
1986). Neurological complications indude polyneuropathy, headache, psychosis, coma, seizures, and
ischaemic infarctions (Bicknell and Holland 1978).
Reversible central vestibulo-auditory dysfunction
has also been described (Benitez et al. 1990).
Differential diagnoses are Meniere's disease with eye
symptoms, syphilitic, chlamydial, or streptomycintreated tuberculous infections as weIl as sarcoidosis
and other auto immune diseases producing vasculitis
(p. 27; Haynes et al. 1980).
How to monitor activity in Cogan's syndrome

Arnold and Gebbers (1994) detected antibodies to


the inner ear by using immunofluorescent labelling
of sections of cadaveric temporal bones with patient
serum. However, no available autoimmune test can
reliably identify the stage of Cogan's syndrome.
Calcified obliterations of bone and soft tissue of the
intralabyrinthine fluid spaces can be visualised in
patients with this syndrome (Majoor et al. 1993;
Casselman et al. 1994). These signal changes are
compatible with bony and soft tissue obliterations of
the endo- and perilymphatic fluid spaces, the presence of acidophilic coagulum, and the hypertrophy
of the stria vascularis. All were found in histopatho-

logical temporal bone studies of patients with


Cogan's syndrome (Wolff et al. 1965; Vollertsen et al.
1986). Casselman et al. (1994) first reported gadolinium enhancement of the cochlea and vestibular
labyrinth in a single case. HeImchen et al. (1998)
assessed the dinical significance of this enhancement for dinical follow-up of the chronic-relapsing
form of Cogan's syndrome. They identified pathological MRI signals in the vestibulum, semicircular
canals, and cochlea during acute exacerbations
which disappeared after the relapse. Moreover, in
chronic forms of Cogan's syndrome with permanent
audiovestibular deficits but no dinical signs of acute
relapse, they found narrowing or obliteration in
parts of the vestibular labyrinth (semicircular canal)
and the cochlea on 3D-CISS images, but no high signal lesions (T \) and no enhancement (Figs
9.13-9.16). Thus, gadolinium-enhanced MRI is useful for differentiating among the different stages of
activity in Cogan's syndrome (HeImchen et al. 1998).
This is of dinical importance, because Cogan's syndrome like other autoimmune inner ear disorders is
treatable by immunosuppressants.

Tumours
Acoustic neurinomas (schwannomas), which generally arise from the vestibular part of the eighth cranial nerve in the internal auditory canal, account for
about 75% of cerebellopontine angle tumours
(Gonzalez-Revilla 1948),5% of which are bilateral
(Fig. 8.3) and pathognomonic for neurofibromatosis
type II (Young et al. 1970). Initially, they present with
slowly progressive hearing loss and tinnitus due to
cochlear nerve compression. Slowly growing
acoustic neurinomas produce such a gradual reduction in vestibular brainstem input from the endorgan on the side of the tumour that the central
compensatory mechanisms are capable of either preventing or minimising the vertigo. In contrast, adequate compensation for the ever-changing reduction
in tonic input of a rapidly growing tumour is not
possible, and vertigo may be prominent, most often
as postural imbalance or disequilibrium. Once the
lesion compresses the brainstem and vestibulocerebellum, central compensation becomes impaired,
and the symptoms of vertigo, oscillopsia, ataxia, postural imbalance, and various ocular motor deficits
(e.g. Brun's nystagmus) decidedly increase. When the
tumours are large, the frequency of central vestibular involvement increases (Berrettini et al. 1996).
Compression of the brainstem or cerebellum is also
indicated by the deviation of the subjectively adjusted

156

Vertigo

Fig.9.14. Acute relapse of Cogan's syndrome.A T,-weighted


2D-FLASH images, axial plane, without contrast. The basal turn of
the left cochlea (= Cl, the vestibule (= V) on both sides, and the
left lateral semicircular canal (LSC) show high signal intensities.
Normally,on T,-weighted images intralabyrinthine fluid is isointense with the cerebrospinal fluid in the internal auditory canal
and cerebellopontine angle. The entry zones of the vestibulocochlear (/ower trace) and facial nerves can be identified at the
brainstem. B T,-weighted 2D-FLASH images, axial plane, after
administration of intravenous gadolinium contrast. Enhancement
is seen in the basal turn of the cochlea (C), in the vestibule (V) on
both sides, and in the left lateral semicircular canal (LSC). The
enhancement is most obvious in the right cochlea, because it
showed no high signal intensities before contrast. Normally, there
is no contrast enhancement in these structures at all. C T,weighted 2D-FLASH images, axial plane, immediately adjacent
slice to (b), after administration of intravenous gadolinium. In
addition to the enhancement in the vestibule (V), the basal turn
of the cochlea (C), and the lateral semicircular canal (LSC), there is
additional enhancement in the posterior semicircular canal (PSC)
on both sides and also in the inferior part of the vestibular nerve
(VN) on both sides. (From Heimchen et al. 1998.)
Fig. 9.13. Cogan's syndrome. T2 -weighted 3D-CISS images,
axial planes of the lateral semicircular canal. A Patient 1: images
during the chronic stage of Cogan's syndrome. Complete occlusion (=1, arrow) and severe narrowing (=2) are shown in the right
lateral semicircular canal. Below, a section of the posterior semicircular canal duct (PSC) can be seen. The vestibulocochlear (=3)
and the facial nerves (FN) can be identified. B Normal subject for
comparison: left lateral semicircular canal is shown (LSC). The
macula (M) is visible at the top of the lateral semicircular canal.
(From Heimchen et al. 1998.)

sense of the visual vertical, ipsiversive to the lesion


(Friedmann 1970). Large tumours also compress
facial and trigeminal nerves and may cause
increased intracranial press ure and hydrocephalus
by obstructing cerebrospinal fluid flow.
MR imaging of the cerebellopontine angle allows
even small intracanalicular tumours (Fig. 9.17) to be
detected and safely removed in toto via a
translabyrinthine or transtemporal approach. The
dassic suboccipital approach is preferred for larger
tumours (Figs 9.18 and 9.19).

Other tumours of the cerebellopontine angle are


meningeomas (Figs. 9.20 and 9.21), which usually
arise dose to the sigmoid and petrosal sinuses
(Nager 1964), and epidermoid cysts. When epidermoid carcinomas enter the middle and the inner
ear, they cause combined audiovestibular loss,
facial palsy, and otorrhea (Lewis 1960; Schuknecht
1974). The tumour is frequentlyvisible in the external auditory canal, and erosion of the temporal
bone is apparent on X-ray examination (Baloh
1984).

Miscellaneous vestibular nerve and labyrinthine disorders

Fig.9.15. Chronic stage of Cogan's syndrome. T2-weighted TSE


images, axial plane. Three sequential slices from rostral to caudal
(A-C),3 mm slice thickness, through the entire labyrinth on the
right (R) and left (L) side are shown. In this sequence, normal
high-signal-intensity fluid, isointense with cerebrospinal fluid,
can be detected in the vestibule (V) and the cochlea (C) on both
sides (white arrows in A-C), but all six semicircular canals are
absent (see Fig. 9.16 for comparison). This indicates soft tissue
obliteration since high-resolution CT did not show any calcification in the semicircular canals. (a): The course of the superior
vestibular nerve (VN) and the facial nerve (FN) can be identified.
The low signal structure between both nerves is the anterior inferior cerebellar artery (AICA). (c) The bony spiral lamina can be
identified in the middle of the cochlea; this helps to differentiate
it from the posterior semicircular canal (see also Fig. 9.16 ). (From
Heimchen et al. 1998.)

Glomus body tumours (Fig. 9.22) along the vagal


or glossopharyngeal nerves are the most common
tumours of the middle ear, but they affect the
labyrinth and caudal cranial nerves more often if
located at the jugular bulb (glomus jugulare) or the
vagus nerve (glomus vagale). Figure 9.23 shows a
diverticle of the jugular vein.
Metastatic carcinoma involving the temporal
bone, e.g. carcinomas of the breast, lung, kidney,

157

Fig.9.16. Normal subject: for comparison with Fig. 9.15, four


sequential slices through the entire labyrinth are shown. T2 weighted TSE images, axial plane. Normal high-signal-intensity
intralabyrinthine fluid can be seen in the cochlea (Cl, the
vestibule, the lateral (arrowheads, LSC), posterior (PSC), and anterior (ASC) semicircular canals on both si des. (From Heimchen et
al. 1998.)

larynx, stomach, prostate, or thyroid gland, invades


the adjacent cranial nerves but rarely destroys the
bony labyrinth (Schuknecht et al. 1968). Meningeal
carcinomatosis (Fig. 8.4 and Fig. 9.24) also affects
audiovestibular function.
Finally, the labyrinth and the eighth nerve can be
damaged by longitudinal fractures of the temporal
bone (Fig. 9.25; see also Chap. 23).

158

Vertigo

Miscellaneous vestibular nerve and labyrinthine disorders

159

Fig.9.18. Left intra-extracanalicular acoustic neurinoma.a Isointense mass lesion in the left cerebellopontine angle (proton-weighted
sequence TR/TE = 2000/28 ms). b Application of intravenous Gd-DTPA causes enhancement of the tumour tissue within internal auditory canal and cerebellopontine angle, a characteristic "pipe-sign" (T,-weighted sequence TR/TE = 500/17 ms) .

...
Fig.9.17. 5chwannoma of the lateral semicircular canal. a Axial projection, T2 -weighted 3D-(155. 5chwannoma can be delineated in
the ampulla of the lateral semicircular canal as a hyperintense lesion with a sharp border surrounded by the hyperintense Iymph. b
Axial projection, T,-weighted 2D-FLA5H. The schwannoma has a higher signal than the surrounding Iymphatic space. c Axial projection, T,-weighted 2D-FLA5H, post Gd-DTPA. Homogeneous enhancement of the schwannoma is shown. (From Jger et al. 1997).

160

Vertigo

b
Fig.9.19. Acoustic neuroma. a Axial projection, T1-weighted 2D-FLASH. An acoustic neuroma (arrowhead) can be delineated from
the internal auditory canal (sm all arrow) into the cerebellopontine angle. b Axial projection, T1-weighted 2D-FLASH, post Gd-DTPA. An
acoustic neuroma (arrowhead) can be delineated from the internal auditory canal (small arrow) into the cerebello-pontine angle.A
homogeneous enhancement and a dural tail sign are seen. (From Jger et al. 1997).

Miscellaneous vestibular nerve and labyrinthine disorders

161

Fig.9.20. Meningeoma of the right cerebellopontine angle and the pyramid apex. a Isointense homogeneous mass (T)-weighted
sequence TR/TE = 500/25 ms). b Intravenous application of Gd-DTPA causes inhomogeneous enhancement, characteristic for
meningeoma (TR/TE = 500/25).

Fig.9.21. Meningeoma.Coronal projection, T)-weighted 2D-FLASH, post Gd-DTPA. Meningeoma (curved arrow) in the cerebellopontine angle. Enhancement of the meningeoma and the dural tail sign (small arrow) is visible. (From Jger et al. 1997).

162

Vertigo

c
Fig.9.22. Left jugular glomus tumour and mastoiditis (MRI).ln the T,-weighted sequence (a: TRfTE = 500/17 ms, axial), a mass lesion
is seen in the left jugular bulb (white arrows).ln addition, pneumatisation is missing in the left mastoid, secondary to inflammation.
After application of the contrast medium (b: TR/TE = 500/17 ms, axial) signal intensity of the tumour is enhanced.ln T2 -weighted
sequence (c: TRfTE =2000/70 ms, axial) mastoiditis presents with high signal intensity.

163

Miscellaneous vestibular nerve and labyrinthine disorders

b
Fig.9.23. Divertide of the jugular vein. Axial (a) and coronal (b) projections, T,-weighted 2D-FLASH, post Gd-DTPA. Diverticulum
jugulare (arrow) in the temporal bone dose to the internal acoustic canal. (From Jger et al. 1997).

164

Vertigo

Fig.9.24. Melanoma with multiple leptomeningeal metastases affecting both eighth nerves with bilateral vestibular and sensorineural hearing 1055. T,-weighted post Gd-DTPA.

Fig.9.25. Longitudinal fracture of the temporal bone (arrow). Axial projection, (T. The fracture runs to the geniculate ganglion of the
facial nerve and then follows the carotic canal. (From Jger et al. 1997).

Miscellaneous vestibular nerve and labyrinthine disorders

References
Arbusow v, Dieterich M, Strupp M, Dreher A, Jger L, Brandt T
(1998a) Herpes zoster neuritis involving superior and inferior
parts of the vestibular nerve causes ocular tilt reaction. Neuroophthalmology 19:17-21
Arbusow V, Strupp M, Dieterich M, Stcker W, Naumann A, Schulz
P, Brandt T (1998b) Serum antibodies against membranous
labyrinth in patients with "idiopathic" bilateral vestibulopathy. J
NeuroI245:132-136
Arnold W, Pfaltz R (1987) Critical evaluation of the immunofluorescence test for identification of serum antibodies against
inner ear tissue. Acta Otolaryngol (Stockh) 103:373-378
Arnold W, Pfaltz R,Altermatt HJ (1985) Evidence of serum antibodies against inner ear tissues in blood of patients with certain sensorineural hearing disorders. Acta Otolaryngol
(Stockh) 99:437-444
Arnold W, Gebbers JO (1994) Serum-Antikrper gegen Korneaund Innenohrgewebe beim Cogan Syndrom. Laryngol Rhinol
OtoI63:428-432
Baloh RW (1984) Dizziness, hearing loss, and tinnitus: the essentials of neurootology. Davis, Philadelphia
Barr B, Lundstrm R (1961) Deafness following maternal rubella.
Acta Otolaryngol (Stoekh) 53:413
Benitez JT, Arsenault MD, Licht JM, Cohen SD, Greenberg RV
(1990) Evidenee of central vestibulo-auditory dysfunetion in
atypical Cogan's syndrome: a case report. Am J Otol 11: 131-134
Berrettini S, Ravecca F, Sellari-Franeeschini S, Brusehini P, Casani
A, Padoleechia R (1996) Aeoustic neuroma: correlations
between morphology and otoneurological manifestations. J
Neurol Sci 144:24-33
Bicknell JM, Holland JV (1978) Neurologie manifest at ions of
Cogan's syndrome. Neurology (Minneap) 28:278-281
Blackley B, Friedman I, Wright I (1967) Herpes zoster auris assoeiated with facial nerve palsy and auditory nerve symptoms. A
case report with histopathological findings. Acta Otolaryngol
(Stockh) 63:533-550
Brevern M von, Lempert T, Bronstein AM, Kocen R (1997)
Selective vestibular damage in neurosarcoidosis. Ann Neurol
42:117-120
Canalis RF (1996) Infections of the ear and temporal bone. In:
Baloh RW, Halmagyi GM (eds) Disorders of the vestibular
system. Oxford University Press, Oxford, pp 340-352
Casselman JW (1994) Magnetic resonance imaging of the inner
ear. Thesis. Universiteit Gent Faculteit Geneeskunde
Casselman JW, Majoor MH, Albers FW (1994) MR of the inner ear
in patients with Cogan's syndrome.AJNR 15:131-138
Casselman JW, Kuhweide R, Dehaene I, Ampe W, Devlies F (1994)
Magnetic res on an ce examination of the inner ear and cerebellopontine angle in patients with vertigo and/or abnormal findings at vestibular testing. Acta Otolaryngol (Stockh) Suppl
513:15-27
Chan CC, Roberge RG, Whiteup SM, Nussenblatt RB (1995) 32
cases of sympathetie ophthalmia. A retrospective study at the
National Eye Institute, Bethesda, MD, from 1982 to 1992. Arch
OphthalmoI1l3:597-600
Cody DTR, Williams HL (1960) Cogan's syndrome. Laryngoscope
70:447-478
Cogan DG (1945) Syndrome of nonsyphilitic interstitial keratitis
and vestibuloauditory symptoms. Arch OphthalmoI33:144-149
Denny-Brown D, Adams RD, Fitzgerald PJ (1944) Pathologie
features of herpes zoster: A note on "genieulate herpes". Areh
Neurol Psychiatry 51:216-231
Eavey RD, Gao YZ, Sehukneeht HF, Gonzales-Pineda M (1985)
Otologie features of baeterial meningitis of ehildhood. J Pediatr
106:402-407
Friedmann G (1970) The judgement of the visual vertical and

165
horizontal with peripheral and central vestibular lesions. Brain
93:313-328
Gemignani G, Berrettini S, Brusehini P, Sellari-Franeesehini S,
Fusari P, Piragine F, Pasero G, Olivieri I (1991) Hearing and
vestibular disturbances in Beh~et's syndrome. Ann Otol Rhinol
Laryngoll00:459-463
Gonzalez-Revilla A (1948) Differential diagnosis of tumours at the
eerebellopontine reeess. Bull Johns Hopkins Hosp 83:187
Goodhill V (1979) Ear diseases, deafness and dizziness. Harper
and Row, Hagerstown, MD
Grimaldi LME, Luzi L, Martino GV, Furlan R, Nemni R, Antonelli
A, Canal N, Pozza G (1993) Bilateral eighth eranial nerve
neuropathy in human immunodefieieney virus infeetion. J
NeuroI240:363-366
Harris J (1987) Experimental autoimmune sensorineural hearing
loss. Laryngoseope 97:63-76
Harris JP, O'Driseoll K (1996) Autoimmune inner ear disease. In:
Baloh RW, Halmagyi GM (ed) Disorders of the vestibular
system. Oxford University Press, Oxford, pp 374-380
Harris JP, Low NC, House WF (1985) Contralateral hearing loss
following inner ear injury: sympathetic cochleolabyrinthitis?
Am J OtoI6:371-377
Husler R, Vibert D, Koralnik IJ, Hirschel B (1991) Neuro-otological manifestations in different stages of HIV infection. Acta
Otolaryngol (Stockh) SuppI481:515-521
Haynes BF, Kaiser-Kupfer MI, Mason P, Faud AS (1980) Cogan's
syndrome: studies in 13 patients, long-term follow-up and a
review of the literature. Medieine 59:426-441
HeImchen C, Jger L, Bttner U, Reiser M, Brandt T (1998) Cogan's
syndrome: High resolution MRI indieators of activity. J Vestib
Res 8:155-167
Hughes G, Kinney S, Barna B, Calabrese L (1984) Practical versus
theoretical management of auto immune inner ear disease.
Laryngoscope 94:758-767
Hunt JR (1908) A further contribution to the herpetic inflammation of the geniculate ganglion. Am J Med Sei 136:226-241
Huygen PLM, Verhagen WIM (1994) Peripheral vestibular and
vestibulo-cochlear dysfunction in hereditary dis orders. J Vestib
Res 4:81-104
Hyden D, dkvist LM, Kylen P (1979) Vestibular symptoms in
mumps deafness. Acta Otolaryngol (Stockh) SuppI360:182-183
Hyden D, Roberg M, dkvist L (1995) Borreliosis as a cause of
sudden deafness and vestibular neuritis in Sweden. Acta
Otolaryngol (Stockh) SuppI520:320-322
Illum P (1972) The Mondini type of cochlear malformation. Arch
Otolaryngol 96:305-311
Ishiyama A, Ishiyama G, Lopez I, Eversole LR, Honrubia V, Baloh
RW (1996) Histopathology of idiopathic chronic recurrent vertigo. Laryngoscope 106:1340-1346
Ishizaki H, Pyykk I, Nozue M (1993) Neuroborreliosis in the
etiology of vestibular neuronitis. Aeta Otolaryngol (Stockh)
Suppl 503:67-69
Jger L, Strupp M, Brandt T, Reiser M (1997) Bildgebung von
Labyrinth und Nervus vestibularis. Nervenarzt 68:443-458
Kobayashi H, Mizukoshi K, Watanabe Y, Nagasaki T, lto M, Aso S
(1991) Otoneurological findings in inner ear syphilis. Acta
Otolaryngol (Stockh) SuppI481:551-555
Komune S, Nogami K, Inoue H, Uemura T (1993) Bilateral
Mondini dysplasia with normal hearing. ORL J
Otorhinolaryngol Relat Spec 55:143-146
Knigsmark BW, Gorlin RJ (1976) Genetic and metabolie
deafness, Saunders, Philadelphia
Lehnhardt E (1958) Pltzliche Hrstrungen auf beiden Seiten
gleichzeitig oder nacheinander aufgetreten. Z Laryngol Rhinol
OtoI37:1-16
Lewis J (1960) Cancer of the ear: Areport of 150 cases.
Laryngoscope 70:551
Longridge MS (1989) Recurrent vestibulopathy: support for a
viral etiology. J Otolaryngol 18:99-100

166
Mafee MF (1993) MRI and CT in the evaluation of acquired and
congenital cholesteatomas of the temporal bone. J Otolaryngol
22:239-248
Majoor MH,Albers FW, Casselman JW (1993) Clinical significance
of magnetic resonance imaging and computed tomography in
Cogan's syndrome. Acta Otolaryngol (Stockh) 113:625-631
Mark AS, Seltzer S, Nelson-Drake J, Chapman JC, Fitzgerald DC,
Gulya AJ (1992) Labyrinthine enhancement on gadoliniumenhanced magnetic resonance imaging in sudden deafness and
vertigo: correlation with audiologic and electronystagmographic studies.Ann Otol Rhinol Laryngol101:459-464
McCabe BF (1979) Autoimmune sensorineural hearing loss. Ann
Otol RhinoI88:585-589
McCabe BF (1989) Autoimmune inner ear disease: therapy. Am J
Otoll0:196-197
McLay K, Maran A (1969) Deafness and Klippel-Feil syndrome. J
Laryngol83:175
Moscicki RA (1994) Immune mediated inner ear disorders. In:
Baloh RW (ed) Neurootology. Balliere Tindall, London, pp
547-563
Moscicki RA, San Martin JE, Quintero CH, Quintero Ch, Rauch
SD, Nadol JB Jr, Bloch KJ (1994) The presence of serum antibody to a 68-kDa inner ear protein identifies a subset of
patients with sensorineural hearing loss and correlates with
disease activity and responsivity to corticosteroid treatment.
JAMA 272:611-616
Nager G (1964) Meningeomas involving the temporal bone: clinical and pathological aspects. Thomas, Springfield
Nance WE, Sweeney A (1975) Genetic factors in deafness in early
life. Otolaryngol Clin North Am 8:19
Pappas DG (1983) Hearing impairments and vestibular abnormalities among children with subclinical cytomegalovirus. Ann
Otol Rhinol Laryngol 92:552-557
Proctor L, Perlman H, Lindsay I, Matz G (1979) Acute vestibular
paralysis in herpes zoster oticus. Ann Otol Rhinol Laryngol
88:403-410
Pyykk I, Levo H, Blomstedt G, Rosenhall U (1997) Sympathetic
cochleolabyrinthitis an occult disease? J Audiol Med 6:24-35
Rosenhall U, Hanner P, Kajiser B (1988) Borrelia infection and
vertigo.Acta Otolaryngol (Stockh) 106:111-116
Rowe-Jones JM, Macallan DC, Sorooshian M (1990) Polyarteritis
nodosa presenting as bilateral sudden onset cochleo-vestibular
failure in a young woman. J Laryngol Otoll04:562-564
Savundra P (1996) Audio-vestibular dysfunction in the sickle cell
syndromes. J Audiol Med 5:167-173
Savundra P, Skacel P, Rudge P (1996) Peripheral vestibulopathy in
sickle cell disease. J Audiol Med 5:61-66

Vertigo
Schuknecht HF (1974) Pathology of the ear. Harvard University
Press, Cambridge Mass
Schuknecht HF (1980) Mondini dysplasia. A clinical and pathological study.Ann Otol Rhinol Laryngol (SuppI65) 85:1-23
Schuknecht HF (1985) Neurolabyrinthitis. Viral infections of the
peripheral auditory and vestibular systems. In: Nomura Y (ed)
Hearing loss and dizziness. Igaku-Shoin, Tokyo, New York, pp
1-15
Schuknecht H, Allam A, Murakami Y (1968) Pathology of secondary malignant tumours of the temporal bone. Ann Otol
Rhinol Laryngol 77:5
Schulz P, Arbusow V, Strupp M, Dieterich M, Sautier W, Brandt T
(1999) Sympathetic contra lateral vestibulopathy after unilateral
zoster oticus. 0 Neurol Neurosurg Psychiatry 66: 672-676)
Seltzer S, Mark AS (1991) Contrast enhancement of the labyrinth
on MR scans in patients with sudden hearing loss and vertigo:
evidence oflabyrinthine disease. Am J NeuroradioI12:13-16
Terjung B, Heimchen C, Samtleben W (1993) Glucocorticoid
monotherapy for Cogan's syndrome? Dtsch Med Wochenschr
118:1231-1235
Tsunoda I, Kanno H, Watanabe M, Shimoji S, Hirayama K, Sumita
H, Yamamoto T (1994) Acute simultaneous bilateral vestibulocochlear impairment in neuro- Behcet's disease: a case report.
Auris Nasus Larynx 21:243-247
Verhagen WIM, Huygen PLM (1994) Central vestibular, vestibuloacoustic and oculomotor dysfunction in hereditary disorders. J
Vestib Res 4: 105-135
Verhagen WIM, Huygen PLM, Horstink NWIM (1987) Familial
congenital vestibular arefiexia. J Neurol Neurosurg Psychiatry
50:933-935
Vollertsen RS, McDonald TI, Younge BR, Banks PM, Stanson AW,
Ilstrup DM (1986) Cogan's syndrome: 18 cases and a review of
the literature. Mayo Clin Proc 61:344-361
Watanabe Y, Aso S, Ohi H, Ishikawa M, Mizukoshi K (1993)
Vestibular nerve dis order in patients suffering from sudden
deafness with vertigo and/or vestibular dysfunction. Acta
Otolaryngol (Stockh) SuppI504:109-111
Wolff D, Bernhard WG, Tsutsumi S, Ross IS, Nussbaum HE (1965)
The pathology of Cogan's syndrome causing profound deafness. Ann Otol Rhinol Laryngol 74:507-520
Young D, Eldridge R, Gardner W (1970) Bilateral acoustic neuroma in a large kindred. JAMA 214:347
Zajtchuk J, Matz G, Lindsay J (1972) Temporal bone pathology in
herpes oticus.Ann Otol Rhinol Laryngol81:331

SECTION C
Central vestibular dis orders

Topographie diagnosis in neurology is frequently


based on lesions along the course of long motor or
sensory pathways. This is well established for the
pyramidal tract and the visual pathways. Likewise,
clinical studies of the differential effects of central
vestibular pathway lesions have increasingly shown
vestibular syndromes to be ace urate indicators for a
topographie diagnosis. Vestibular pathways run
from the eighth nerve and the vestibular nuclei
through ascending fibres, such as the ipsilateral or
contralateral medial longitudinal fasciculus (MLF),
the brachium conjunctivum, or the ventral tegmental tract to the oculomotor nuclei, the supranuclear
integration centres in the rostral midbrain, and the
vestibular thalamic subnuclei. From there they reach
several cortex areas through the thalamic projection.
Another relevant ascending projection reaches the
cortex from vestibular nuclei via vestibular cerebellum structures, in particular the fastigial nucleus.
In the majority of cases, central vestibular vertigo
syndromes are caused by dysfunction or a deficit of
sensory input induced by alesion. In a small proportion of cases they are due to pathological excitation
of various structures, extending from the peripheral
vestibular organ to the vestibular cortex (see paroxysmal vertigo, p. 29). Since peripheral vestibular disorders are always characterised by a combination of
perceptual, ocular motor, and postural signs and
symptoms, central vestibular disorders may manifest
as "a complete syndrome" or with only single components. The ocular motor aspect, for example, predominates in the syndromes of upbeat or downbeat
nystagmus. Lateral falls may occur without vertigo
in vestibular thalamic lesions (thalamic astasia) or
as lateropulsion in Wallenberg's syndrome.

all patients who complain of dizziness (p. 5; Leigh


and Brandt 1993) and are apart of the examination
of the unconscious patient. There is evidence for a
useful clinical classification of central vestibular syndromes according to the three major planes of
action of the VOR (Fig. c.l): yaw, roll, and pitch
(Brandt 1991; Brandt and Dieterich 1994,1995).
The plane-specific vestibular syndromes are
determined by ocular motor, postural, and perceptual signs (Fig. C.2):

Yaw plane signs are horizontal nystagmus, past


pointing, rotational and lateral body falls, horizontal deviation of perceived straight -ahead.
Roll plane signs are torsional nystagmus, skew
deviation, ocular torsion, tilts of head, body, and
perceived vertical.
Pitch plane signs are upbeat/downbeat nystagmus, forward/backward tilts and falls, vertical
deviation of perceived straight-ahead.
YAW

Clinical classification of central


vestibular disorders
The "elementary" neuronal network of the vestibular
system is the di - or trisynaptic vestibulo-ocular
reflex (VOR). VOR properties are routinely tested in

Fig. c'1. Schematic representation of the three major planes of


action,of the vestibulo-ocular reflex. Horizontal rotation about
the vertical z axis = yaw; vertical rotation about the binaural y
axis = pitch; vertical rotation about the x axis ("li ne of sight") =
roll.

169

170

Vertigo

The thus-defined VOR syndromes allow for a preeise topographie diagnosis of brainstem lesions as to
their level and side (Fig. C.3; Brandt and Dieterieh

1994,1995).

A tone imbalance in yaw indieates lesions of the


lateral medulla including the root entry zone of
the eighth nerve and/or the vestibular nudei.
A tone imbalance in roll indieates unilateral

torsional nystagmus
skew deviation
ocular torsion (skew tors io n)
head and body (roll) tilt
and fa lls (ipsi or contra)
lill of perceived vertical
(ipsi or contra)

vertical nystagmus (up I down)


up/down ocular deviation
fore - aft body sway and falls
upbeat:downward deviation
of subjective straight ahead
downbeat :upward deviation
of subjective straight ahead

horizontal nystagmus
horizontal ocular deviation
caloric hyporesponsiveness
(ipsilateral)
horizontal body rotation.
lateral falls (ipsiversive)
past- pointing
horizontal deviation of
subjective straight ahead

Fig. C.2. Topographic diagnosis of vestibular syndromes in roll, pitch, and yaw planes: schematic presentation of the distinct areas
within the brainstem and vestibulocerebellum (frontal and sagittal 'views) in wh ich alesion induces a vestibulo-ocular tone imbalance
in roll, pitch, or yaw plane. Typical ocular motor, postural, andperceptual signs are torsional, vertical (up/downbeatl. or horizontal
nystagmus. A tone imbalance in roll indicates unilateral "graviceptive" pathway lesions from the medial or superior vestibular nuclei
(inducing ipsiversive signs), crossing midline to the contra lateral MLF and the rostral integration centres for vertical and torsional eye
movements, the INC and riMLF (inducing contraversive signs). A tone imbalance in pitch indicates paramedian bilateral brainstem
lesions at pontomesencephalic or pontomedullary level, the brachium conjunctivum, or the flocculi.lt is striking that pontomedullary
lesions may induce either upbeat or down beat nystagmus or transitions between the two, whereas binocular flocculus lesions result
only in down beat nystagmus and a pontomesencephalic lesion only in upbeat nystagmus. A tone imbalance in yaw indicates a unilateral pontomedullary lesion involving the medial and superior vestibular nucleus. This area overlaps with roll and pitch function of the
VOR, which explains the frequency of mixed vestibular syndromes in more than one plane. (riMLF = rostral interstitial nucleus of the
medial longitudinal fasciculus, INC = interstitial nucleus of Cajal, 111 = oculomotor nucleus, IV = trochlear nucleus, VI = abducens
nucleus, VIII = vestibular nucleus.) (From Brandt and Dieterich 1995.)

171

Central Vestibular Disorders

lesions (ipsiversive at pontomedullary level, contraversive at pontomesencephalic level).


A tone imbalance in pitch indicates bilateral
(paramedian) lesions or bilateral dysfunction of
the flocculus.

Consequently, discussion of central vestibular disorders in Chapters 10-12 will follow this classification:

Vestibular dis orders in (frontal) roll plane


Chapter 10
Vestibular disorders in (sagittal) pitch plane
Chapter 11
Vestibular dis orders in (horizontal) yaw plane
Chapter 12

It is hypothesised that signal processing of the


VOR in roll and pitch is conveyed by the same rather

INe
D

yaw

[2J up
> '1 h
[5] down pi C

Mesencephalon

roll

-+--- Pons

Flocculus
- . - - --

Medulla

Fig. C3. Vestibular syndromes in roll, piteh and yaw planes:


eritieal areas are sehematieally represented based on our eurrent
knowledge of vestibular and oeular motor struetures and pathways, alesion of which causes a vestibular tone imbalance in one
of the three major planes of action. The mere clinieal sign of a vertieal, torsional, or horizontal nystagmus - if eentral-vestibularallows a topographie diagnosis of the lesion, although the partieular vestibular struetures involved are still under diseussion.
Whereas a vestibular tone imbalanee in the roll plane indieates
unilateral brainstem lesions (a erossing in the pons), vertieal nystagmus indieates bilaterallesions. Two separate eausative loei are
known for upbeat nystagmus: medullary or pontomeseneephalic.
Downbeat nystagmus indieates abilateral para median lesion of
the eommissural fibres between the vestibular nuclei or a bilateral floeeulus lesion. Horizontal nystagmus indieates unilateral pontomedullary lesions involving the vestibular nuclei. The
differentiation of vestibular oeular motor signs aeeording to the
three major planes of action of the VOR and their mapping to distinet and separate areas in the brainstem are helpful for topographie diagnosis and for avoiding ineorreet assignment of
clinieal signs to brainstem lesions identified with imaging teehniques (lNC = interstitial nucleus of Cajal, MLF = mediallongitudinal faseiculus, VN = vestibular nucleus). (From Brandt and
Dieterieh 1994).

than separate ascending pathways in the medial longitudinal fasciculus and the brachium conjunctivum. A unilaterallesion (or stimulation) of these
"graviceptive" pathways (which transduce input
from vertical semicircular canals and otoliths)
affects function in roll, whereas bilaterallesions (or
stimulation) affects function in pitch. Thus, the
vestibular system is able to change its functional
plane of action from roll to pitch by switching from
a unilateral to a bilateral mode of operation (Fig. C.4;
Brandt and Dieterich 1995).
Clinically this means that "bilateral syndromes in
roll (skew torsion)" appear as one syndrome in pitch
(upbeat or downbeat nystagmus). Pure syndromes
in yaw are rare, since the small causative area covering the medial and superior vestibular nucleus is not
only adjacent to but overlapped by the structures
also subserving roll and pitch function. Alesion frequently results in mixed (e.g. torsional and horizontal) nystagmus. The lesional sites of yaw syndromes
are restricted to the pontomedullary level because of
the short distance between the vestibular nuclei and
the integration centre for horizontal eye movements
in the paramedian pontine reticular formation.
Syndromes in roll and pitch, however, may arise
from brainstem lesions located in an area extending
from the medulla to the mesencephalon, an area corresponding to the large distance between the
vestibular nuclei and the integration centres for vertical and torsional eye movements in the rostral
midbrain. Whereas vestibular tone imbalances in
pitch - which involve bilateral pathways - may occur
with various intoxications or metabolic dis orders,
this is an unusual aetiology for tone imbalances in
yaw or roll - which involve vestibular pathways
unilaterally.
Some vestibular disorders are characterised by a
simultaneously peripher al and central vestibular
involvement. Examples are large acoustic neurinomas (p. 155), infarctions of the anterior inferior cerebellar artery (p. 308), head trauma (p. 347), and
syndromes induced by alcohol intoxication (p. 399).
Others may affect the vestibular nerve root in the
brainstem, where the transition between the peripheral and central nervous system has been defined as
the Redlich-Oberstein zone (lacunar infarction, focal
demyelination in MS, p. 241).
Cortical vestibular syndromes include vestibular
seizures (vestibular epilepsy,p. 233) and dysfunction
with tilt of the perceived vertical, lateropulsion
(p. 192), rarely rotation al vertigo (p. 318). There is
no primary vestibular cortex (p. 219), but cortical
vestibular function is imbedded in a network of multisensory visual-vestibular-somatosensory functions
and distributed over several separate and distinct
areas in the temporoparietal region.

Vertigo

172

LEFT LABYRINTH

RIGHT LABYR INTH

0J~~~
.)J CD
+. 0 ......--......

PONS

MESENCEPHALON

CD
CD
0)

I'"

I'"

~~~t
t~~t

t~Li>~

The parieto-insular vestibular cortex (Guldin and


Grsser 1996; p. 220) seems to act as a kind of main
integration centre. Dysfunction of this multisensory
and sensorimotor cortex for spatial orientation and
self-motion perception may be involved in spatial
hemineglect (p. 224) and rare paroxysmal room-tilt
illusions (p. 224). Visual-vestibular interaction in
self-motion perception is obviously based on reciprocal inhibitory interaction (p. 225), a mechanism
that makes perception of self-motion more robust
and largely insensitive to visual-vestibular mismatches.
Most central vertigo syndromes have a specific
locus (Table C.!) but not a specific aetiology. The
aetiology may, for example, be vascular, autoimmunological (e.g. in MS) inftammatory, neo-plastic,
toxic, or traumatic.

ROLL

PITCH

ROLL

Ffg. C.4. Hypothetical explanation of how the major plane of


action of the VOR may change from roll to pitch depending on
the unilateral or bilateral operational mode of common "graviceptive" pathways: unilateral stimulation (or lesion) causes an
imbalance in roll, whereas bilateral stimulations (or lesions) cause
an imbalance in pitch, because the two (counterdirected) torsional components cancel each other. Schematic representation of
crossed excitatory (posterior semicircular canal (pe)) pathways,
wh ich convey a mixed vertical and torsional ocular motor
impulse to both eyes (both eyes downward, right eye incyclotorsion, left eye excyclotorsion). A unilaterallesion of the right pathway (1) caudal to the midline crossing results in tone imbalance
in roll with skew torsion (unaffected pathway input drives the
right eye downward and counterclockwise, whereas disruption of
the input of the affected pathway causes the left eye to move
upward and counterclockwise). A bilaterallesion (2) results in a
transition of the plane of action from roll to pitch because torsional components cancel each other, and the 1055 of downward
impulse leads to an upward deviation of the eyes. Upward deviation with downward refixation saccades manifests as down beat
nystagmus. A unilaterallesion rostral to the pathway crossing (3)
causes a directional pendant to lesion (1); the skew torsion is now
contraversive. Theoretically, VOR in roll and pitch can be mediated
bya common rather than aseparate pathway system integrating
convergent information from all four vertical semicircular canals
and otoliths. (From Brandt and Dieterich 1995).

References
Brandt Th (1991) Man in motion. Historical and clinieal aspeets of
vestibular funetion. Brain 114:2159-2174
Brandt T, Dieterieh M (1994) Vestibular syndromes in the roll
plane: topographie diagnosis from brainstem to cortex. Ann
NeuroI36:337-347
Brandt T, Dieterieh M (1995) Central vestibular syndromes in the
roll, piteh, and yaw planes: topographie diagnosis of brainstem
disorders. Neuroophthalmology 15:291-303
Guldin W, Grsser 0- I (1996) The anatomy of the vestibular eortiees of primates. In: M Collard, M Jeannerod, Y Christen (eds)
Le cortex vestibulaire. Ipsen, Boulogne, pp 17-26
Leigh J, Brandt T (1993) Areevaluation of the vestibulo-oeular
reflex: new ideas of its purpose, properties, neural substrate,
and disorders. Neurology 43: 1288-1295

173

Central Vestibular Disorders


Table C.l. Central vestibular syndromes
Site

- - - - _ ..

__ ..

Syndrome

_-~-_._----'--------

Vestibular cortex (multisensory)

-_.

__ .._ -

Vestibular epilepsy
Volvular epilepsy
Non-epileptic cortical vertigo
Spatial hemineglect (contraversive)
Transient room-tilt illusions

Tilt of perceived vertical with body


lateropulsion (mostly contraversive)

Mechanism/aetiology

-"'--------------

Vestibular seizures are auras (simple or complex partial


multisensory seizures)
Sensorimotor "vestibular" rotatory seizures
with walking in small circles
Rare rotatory vertigo in acute lesions of the parieto-insular
vestibular cortex
Multisensory horizontal deviation of spatial attention with parietal
or frontal cortex lesions (non-dominant hemisphere)
Paroxysmal or transient mismatch of visual- and vestibular 3-D
spatial coordinate maps in vestibular brainstem, parietal, or frontal
cortex lesions
Vestibular tone imbalance in roll with acute lesions of
the parieto-insular vestibular cortex

Thalamus

Thalamic astasia
Tilt of perceived vertical (ipsiversive or
contraversive) with body lateropulsion

Dorsolateral vestibular thalamic lesions


Vestibular tone imbalance in roll

Mesodiencephalic brainstem

Dcular tilt reaction (contraversive;


ipsiversive if paroxysmal)
Torsional nystagmus (ipsiversive or
contraversive)

Vestibular tone imbalance in roll (integrator-DTR with


INC lesions)
Ipsiversive in INC lesions
Contraversive in riMLF lesions

Mesencephalic brainstem

Skew torsion (contraversive)


Upbeat nystagmus

Vestibular tone imbalance in roll with MLF lesions


Vestibular tone imbalance in pitch in bilateral brachium
conjunctivum lesions

Ponto-medullary brainstem

Tilt of perceived vertical, lateropulsion,


ocular tilt reaction
Pseudo "vestibular neuritis"
Downbeat nystagmus
Transient room-tilt illusion
Paroxysmal room-tilt illusion in MS
Paroxysmal dysarthria/ataxia in MS
Paroxysmal vertigo evoked by lateral gaze

Vestibular tone imbalance in roll with medial and/or


superior vestibular nuclei lesions
Lacunar infarction or MS plaque at the root entry zone of the
eighth nerve
Vestibular tone imbalance in pitch
Acute severe vestibular tone imbalance in roll or pitch
Transversally spreading ephaptic axonal activity
Transversally spreading ephaptic axonal activation
Vestibular nuclei lesion?

Medulla

Upbeat nystagmus

Vestibular tone imbalance in pitch? (nucleus prepositus hypoglossi)

Vestibular cerebellum

Downbeat nystagmus

Vestibular tone imbalance in pitch caused by bilateral flocculus


lesions (disinhibition)
Disinhibited otolith-canal interaction in nodulus lesions?
EA 1=autosomally dominant inherited potassium
channelopathy
EA2=autosomally dominant inherited calcium channelopathy
Viral infection of cerebellum
Viral infection of cerebellum

Positional down beat nystagmus


Familial episodic ataxia (EA 1 with myokymia
and EA2 with vertigo)
Encephalitis with predominant vertigo
Epidemic vertigo

Vestibular disorders in (frontal)


roll plane

The clinieal signs, both pereeptual and motor, of a


vestibular tone imbalanee in the roll plane include
ocular tilt reaetion (OTR), oeular torsion, skew deviation, and tilts of the pereeived (subjeetive) visual
vertieal (SVV). Complete OTR or skew torsion without head tilt indicates either a unilateral peripheral
deficit of otolith and vertical semicireular eanal
input or a unilateral lesion of "graviceptive" brainstern pathways from the vestibular nuclei (erossing
midline at the pontine level) to the interstitial nucleus
of Cajal (I C) in the rostral midbrain. VV tilts are
the most sensitive clinieal sign of an aeute unilateral
brainstem infaretion and signify a vestibular tone
imbalanee in roll. They oeeur with peripheral or eentral vestibular lesions from the labyrinth to the
vestibular cortex.
All tilt effeets - pereeptual, oeular motor, and postural - are ipsiversive (ipsilateral eye undermost)
when unilateral peripheral or pontomedullary
lesions oeeur below the erossing of the gravieeptive
pathways. They are eontraversive (contralateral eye
undermost) in eases of unilateral pontomeseneephalie brainstem lesions, and indieate involvement
of the medial longitudinal faseieulus (MLF) or the
rostral midbrain integration centres of the VOR in
roll and piteh planes (INC; riMLF). Vestibular disorders in the roll plane may be distinguished by
the predominant pereeptual or motor signs and
symptoms.
It is also possible to relate the clinieal syndrome
to a partieular brainstem level, the lesioned vestibular struetures, and the involved reflex (VOR) or integration (INC) meehanisffi. Three types of vestibular
disorders in roll ean be thus deseribed (see also
p. 181):
the "ascending" type of VOR-OTR with pontomedullary lesions of the medial or superior
vestibular nucleus or gravieeptive pathways
whieh subserve the VOR in the roll plane,
2. skew torsion without head tilt, with unilateral
lesions of the pontomeseneephalie "graviceptive"
pathways, and
1.

175

3. the "descending" type of integrator-OTR with


lesions of the rostral midbrain integration eentres for eye-head co ordination in the roll and
piteh planes.
The "aseending, reflexive" type (lateral medulla)
of OTR simply refteets a tone im balance of the VOR,
whereas the "deseending integration" type (rostral
midbrain tegmentum) integrates signals of eye-head
velocity with those of holding position and involves
eortieal eontrol of the fundamental pattern of OTR
du ring aetive loeomotion. The (pontomeseneephalie) "skew torsion" type results from lesions of
aseending "graviceptive" pathways but does not
eause head tilt, beeause vestibulospinal traets are not
involved. Unilateral infarctions of the brainstem (see
Chap. 19) cause dysfunetion predominantly in the
roll plane, less frequently in the yaw and piteh
planes. Dysfunction in the pitch plane requires bilateral vestibular dysfunetion.
Unilaterallesions of vestibular struetures rostral
to the INC typically manifest as deviations of pereeived vertieal without coneurrent eye-head tilt.
OTR in unilateral paramedian thalamie infaretions
indicates simultaneous isehaemia of the paramedian
rostral midbrain including the INC. Unilateral
lesions of the posterolateral thalamus can eause
thalamie astasia and moderate ipsiversive or eontraversive SVV tilts, thereby indicating involvement of
the vestibular thalamic subnuclei. Unilaterallesions
of the parietoinsular vestibular cortex eause moderate, mostly eontraversive SVV tilts. An SVV tilt
found with monoeular but not with binocular viewing is typical for a troehlear or oculomotor palsy
rather than a supranuclear gravieeptive brainstem
lesion.
"Gravieeptive" input converging from otoliths and
vertieal semieireular eanals subserves vestibular
funetion in the roll plane. A unilateraliesion of these
pathways may result in statie and dynamie effeets,
e.g. torsional nystagmus. Several distinet and
separate lesions have been associated with torsional
nystagmus: lesions of the vestibular nuclei, the lateral

Vertigo

176

medulla, and in rare cases the MLF, the I C, and the


rostral interstitial nucleus of the MLF (riMLF). Fast
phases of torsional nystagmus are contraversive in
pontomedullary lesions and ipsiversive in paramedian pontine and pontomesencephalic lesions. Rarely
they may be ipsiversive in circumscribed lesions of
the riMLE
The spontaneous course of vestibular disorders in
the roll plane is characterised bya gradual recovery,
probably based on a complex compensation of the
VOR tone imbalance. The functional significance of
central compensation of a central vestibular tone
imbalance (due to a unilaterallesion) is similar to
central compensation of a peripheral vestibular loss,
as seen, e.g. in vestibular neuritis (see Chap.4; p. 67).

The clinical syndrome


The "graviceptive" input from the otoliths converges
with that from the vertical semicircular canals at the
level of the vestibular nuclei (Angelaki et al. 1993)
and the ocular motor nuclei (Baker et al. 1973;
Schwindt et al. 1973) to subserve static and dynamic
vestibular function in pitch (up and down in the
sagittal plane) and roll (lateral tilt in the frontal
plane). In the "normal" position in the roll plane, the
subjective visual vertical (SVV) is aligned with the
gravitational vertical, and the axes of the eyes and
the head are horizontal and directed straight ahead.
Signs and symptoms of a vestibular dysfunction
in the roll plane can be derived from the deviations
from normal function. A lesion-induced vestibular
tone imbalance should result in a syndrome consisting of a perceptual tilt (SVV), head and body tilt
(OTR, lateropulsion), vertical misalignment of the
visual axes (skew deviation), and ocular torsion (Fig.
10.1). This has been demonstrated in animal experiments by unilateral stimulation of the utricular
nerve (cat: Suzuki et aL 1969), the utricular macula
(guinea-pig: Curthoys 1987), and vertical canal
nerves (cat: Cohen et al. 1964; Tokumasu et al. 1971),
which resuIted in either a complete ocular tiIt reaction (OTR), i.e. the triad of head tiIt, skew deviation,
and ocular torsion, or in its single components. In
humans, inadvertent damage to one utricle
(Halmagyi et al. 1979) or inappropriate stimulation
of the otoliths in patients with a Tullio phenomenon
(Dieterich et al. 1989; p. 108) also caused OTR.
Not only the complete synkinesis of OTR but also
one of its components (postural or perceived tiIt,
ocular torsion, or skew deviation) indicates a dysfunction in the roll plane. Skew deviation appears
primarily as an ocular motor misalignment of the

visual axes in the vertical plane, but its regular association with ocular torsion and SVV tiIt (Brandt and
Dieterich 1993) makes a vestibular roll plane dysfunction most likely.
We can establish clinical rules, which help our
basic understanding and diagnostic routine,
although exceptions may prove them inaccurate. The
only difference between the ocular skew-torsion sign
and OTR is head tiIt. Whereas skew deviation does
not manifest without ocular torsion (exceptions will
probably be found in the future), monocular or
binocular torsion is frequently seen without concurre nt skew deviation. Finally, perceived vertical may
be tiIted with or without concurrent skew deviation,
ocular torsion, or head tiIt (Fig. 10.1, Table 10.1).
There is convincing evidence that all following
signs and symptoms refiect vestibular dysfunction
in the (frontal) roll plane:

ocular tilt re action (OTR)


skew deviation (skew-torsion sign)
spontaneous torsion al nystagmus
tonic ocular torsion (monocular or binocular), if
not caused by infranuclear ocular motor disorders
tilt of perceived visual vertical (SVV) (with
binocular viewing)
body lateropulsion

Ocular motor or postural tilts as weIl as misadjustments of subjective vertical point in the same
direction, either clockwise or counterclockwise (as
seen from the viewpoint of the examiner). The direction of all tiIts is reversed if pathological excitation
of unilateral "graviceptive" pathways is the cause of
vestibular tone imbalance in roll rather than a
lesional input deficit. The combination of static and
dynamic signs is not surprising if one considers the
functional cooperation of otoliths and vertical semicircular canals due to their neuronal convergence
within "graviceptive" pathways. The above-listed
signs and symptoms may be found in combination
or as single components at all brainstem levels. A
systematic study of 111 patients with acute unilateral
brainstem infarctions revealed that pathological tiIts
of SVV (94%) and ocular torsion (83%) are the most
sensitive signs (Dieterich and Brandt 1993a). Skew
deviation was found in one-third and a complete
OTR in one-fifth of these patients (see Table 10.1,
Fig.1O.2).
Clinical evaluation of vestibular function in roll
therefore includes psychophysical adjustments of
the SVV, determination of the vertical divergence of
the visual axes by means of prisms, and determination of ocular torsion by means of fundus
photographs (for methods, see Dieterich and Brandt
1993a).

177

Vestibular disorders in (frontal) roll plane

binocular

left

objective vertical

RE
b
Fig.10.1. a Schematic drawing of ocular tilt reaction (OTR) to the right with rightward head tilt, skew deviation (right eye undermost), and binocular torsion. b Patient with a left paramedian thalamic and interstitial nucleus of Cajal infarction presenting with a
complete OTR to the right. Note: Skew deviation of 10 degrees shown in the fundus photographs (top). ( Adjustments of the subjective visual vertical (SW) under binocular (top) or monocular viewing conditions (means and standard deviations). The direction of SW
tilt corresponds to head tilt. RE = right eye; LE = left eye. (From Brandt and Dieterich 1994.)

Vertigo

178

Table 10.1. Frequency of subjective visual vertical tilt, skew deviation, ocular torsion, and ocular tilt reaction in acute unilateral brainstem
and thalamic infarctions
Patients (n)

Lesion

SVVtilt (%)

Ocular torsion (%)


Monocular
Binocular

Skew(%)

OTR(%)
-

Mesodiencephalic
Para median thalamic
Posterolateral thalamic
Anterior polar thalamic
Mesencephalic
Pontomesencephalic
Pontine
Pontomedullary
Medullary (Wallenberg's syndrome)
Total

14
17
4

64
65
0

16
12
34
13
36

94
92
91
100
94
94

111

~----

29'
13 b
0

43'
20 b
0

57
0
0

57
0
0

54
64
47
60
27
47

38
18
33
20

37.5
25
26.5
23
44
31

25
25
12
7.7
33

55
36

'Additional third nerve palsy.


bSlight torsion of about 2.8.
SVV = subjective visual vertical; OTR = ocular tilt reaction.
From Brandt and Dieterich (1994).

2.

Patients with acute uni lateral


brainstem lesions 100%

3.

Fig.l0.2. Frequeney (%) of different signs and symptoms of


vestibular dysfunetion in the roll plane as determined in a total of
111 patients with aeute unilateral brainstem infaretions.
Subjeetive visual vertical (SVV) tilts are the most sensitive sign,
followed by monoeular or binoeular torsion. They all represent
statie rather than dynamic (torsional nystagmus) signs of gravieeptive pathway lesions; otolithie input is the most important of
these. OTR = oeular tilt reaetion; SD = skew deviation; OT = oeular
torsion. (From Brandt and Dieterieh 1994.)

4.

Topographie diagnostic rules


A crossing of graviceptive pathways has been identified in clinical studies of the differential effects of
unilateral brainstem infarctions (Chap. 19; p. 307) on
vestibular function in the three major planes of
action. These current clinical data support the following preliminary topographie diagnostic rules
based on vestibular signs and symptoms in roll (Fig.
10.3; Brandt and Dieterich 1994,1995):
1.

The fundamental pattern of eye-head tilt in roll


- either complete ocular tilt reaction (OTR) or

5.

6.

skew torsion without head tilt - indicates a unilateral peripheral deficit of otolith and vertical
canal input or a unilaterallesion of"graviceptive" brainstem pathways from the vestibular
nuclei (crossing midline at lower pontine level)
to the interstitial nucleus of Cajal (INC) in the
rostral midbrain.
Tilts of perceived vertical (SVV), resulting from
peripheral or central vestibular lesions from
the labyrinth to the vestibular cortex, are the
most sensitive sign of a vestibular tone imbalan ce in roll (Dieterich and Brandt 1993a).
All tilt effects - perceptual, ocular motor, and
postural - are ipsiversive (ipsilateral eye lowermost) and due to unilateral peripheral or pontomedullary lesions below the crossing of the
graviceptive pathways. They indicate involvement of the labyrinth, vestibular nerve, or
medial and/or superior vestibular nuclei; the
latter are mainly supplied by the vertebral
artery (Dieterich and Brandt 1992).
All tilt effects in unilateral pontomesencephalic
brainstem lesions are contraversive (contralateral eye lowermost) and indicate involvement of
the medial longitudinal fasciculus (MLF) (paramedian arteries arising from basilar artery) or
INC and riMLF (paramedian superior mesencephalic arteries arising from the basilar
artery) (Brandt and Dieterich 1993; Dieterich
and Brandt 1993b).
OTR with unilateral (ponto )medullary lesions
(vestibular nuclei) indicates the "ascending"
(reflexive) type of a tone imbalance of the VOR
in roll (p. 181).
OTR due to rostral midbrain lesions (INC)
reflects the "descending" type of tone imbalance

179

Vestibular disorders in (frontal) roll plane

7.

8.

9.

10.

11.

12.

13.

involving the neural integration cent re for eyehead co ordination in roll (p. 181).
Skew deviation is always combined with ocular
torsion, i.e. skew-torsion sign (Brandt and
Dieterich 1993). It manifests without head tilt if
ascending pontomesencephalic "graviceptive"
pathways are affected rostral to the downward
branching of the vestibulospinal tract.
Unilaterallesions of ascending vestibular pathways rostral to the INC typically manifest with
deviations of perceived vertical without concurrent eye-head tilt (Dieterich and Brandt 1993b).
OTR in unilateral paramedian thalamic infarctions (paramedian thalamic arteries from basilar artery) indicates simultaneous ischaemia of
the paramedian rostral midbrain including the
INC (Dieterich and Brandt 1993b).
Unilaterallesions of the posterolateral thalamus can cause thalamic astasia and moderate
ipsiversive or contraversive SVV tiIts, thereby
indicating involvement of the vestibular thalamic subnuclei (thalamogeniculate arteries )
(Dieterich and Brandt 1993b).
Unilateral lesions of the parieto-insular
vestibular cortex (PIVC) cause moderate,
mostly contraversive SVV tiIts (temporal
branches of the middle cerebral artery or deep
perforators) (Brandt et al. 1994) and "cortical
lateropulsion" .
Tilt effects caused by paroxysmal activation of
"graviceptive" pathways point in the opposite
direction of those caused by lesional inhibition,
such as unilateral infarction (Halmagyi et al.
1990; Lueck et al. 1991; Rabinovitch et al. 1977;
Hedges and Hoyt 1982).
An SVV tilt found with monocular but not with
binocular viewing is typical for a trochlear or
oculomotor palsy rather than a supranuclear
"graviceptive" brainstem lesion (Dieterich and
Brandt 1993c).

Thus, all clinical signs of vestibular dysfunction in


roll can be helpful when determining not only the
level but also the side of the brainstem lesion. If the
level of damage is known from the clinical syndrome, the vestibular syndrome indicates the more
severely affected side. Conversely, if the side of damage is clear from the clinical syndrome, the direction
of OTR, skew deviation, and SVV tiIt indicates the
level on the brainstem. The diagnostic topographic
value of these signs is similar to those of cranial
nerve lesions, but is more sensitive. The effects discussed can be attributed to a lesional vestibular tone
imbalance caused by decreased resting activity of
the "graviceptive" pathways or INC neurons, whereas

Mesencephalon
Pons

Medulla

_._ . utriCI~~

~
~ .\ ~

- vertical
semicircular
canals

"""""""

V,

Fig. 10.3. Vestibular syndromes in roll plane: Graviceptive


pathways from otoliths and vertical semicircular canals mediating vestibular function in roll plane. The projections from the
otoliths and the vertical semicircular canals to the ocular motor
nuclei (trochlear nucleus IV, oculomotor nucleus 111, abducens
nucleus VI), the supranuclear centres of the INe, and the rostral
interstitial nucleus of the MLF (riMLF) are shown. They subserve
VOR in three planes. The VOR is part of a more complex vestibular
reaction, which also involves vestibulospinal connections via the
medial and lateral vestibulospinal tracts for head and body posture control. Furthermore, connections to the assumed vestibular
cortex (areas 2v and 3a and the parieto-insular vestibular cortex,
PIVC) via the vestibular nuclei of the thalamus (Vim, Vce) are
depicted. "Graviceptive" vestibular pathways for the roll plane
cross at the pontine level. OTR (skew torsion, head tilt, and tilt of
perceived vertical, SW) is depicted schematically on the right in
relation to the level of the lesion: ipsiversive OTR with peripheral
and pontomedullary lesions; contraversive OTR with pontomesencephalic lesions.ln vestibular thalamic lesions, the tilts of SVV
may be contraversive or ipsiversive; in vestibular cortex lesions
they are preferably contraversive.OTR is not induced by supratentorial lesions above the level of INe. (From Brandt and
Dieterich 1995.)

paroxysmal effects in the opposite direction have


been ascribed to transiently increased activity or
electrical stimulation.

Ocular tilt reaction (OTR)


OTR is an oculocephalic response consisting of ipsilateral head tiIt, skew deviation, and ocular torsion. It

Vertigo

180

was first clearly delineated by Westheimer and Blair


(1975 a,b), who applied electrical stimulation to the
monkey rostral midbrain tegmentum in the region
of the INe. Earlier stimulation experiments of the
INC, the pretectal area, and the zona incerta of the
cat evoked rotatory movements of the head in the
frontal plane (Bartorelli 1942) and conjugated
cyclorotation of the eyes when the head was held
fixed (Hess et al. 1946). By the 1940s, Szentagothai
(1943) considered the region of the interstitial nucleus
of Cajal (INC) to be a coordination centre of rotatory
eye movements. Hyde and Toczek (1962) were able to
show that rotational movements of the eyes and the
head, evoked by stimulation of the zona incerta of
the cat, depend on the functional integrity of ipsilateral INe. Cyclorotation was significantly decreased
or abolished after destruction in or medial to the
INC; head tilts in the frontal plane were significantly
decreased after alesion posteromedial to the INe. A
decade later, Sano et al. (1972) reported that stereotactic electrical stimulation of the INC or adjacent
areas of the medial longitudinal fasciculus in
humans induced either head extension (retroflexion)
or ipsilateral head tilt (in cases of torticollis). When
stimulating the INC in one of their patients (history
of post-traumatic see-saw nystagmus), they activated
see-saw nystagmus accompanied by head tilt (probably
a complete OTR, unknown at that time). Electrocauterisation of the area abolished these signs. When
the area adjacent and caudal to the INC was stimulated, only head tilt occurred.
Only three human cases were reported during the
decade following the original description of OTR by
Westheimer and Blair in 1975. A paroxysmal OTR
was reported by Rabinovitch et al. (1977) in a patient
with multiple sclerosis and by Hedges and Hoyt
(1982) in a patient with a brainstem abscess in the
region of the zona incerta. Halmagyi et al. (1979)
documented a reversible tonic OTR in a patient with

a partial iatrogenic lesion of the utricle following


stapedectomy. The first non-paroxysmal tonic central OTR persisting for years in patients with rostral
midbrain lesions (Fig. 10.4) and the first patients
with OTR due to lateral medullary brainstem lesions
were described by Brandt and Dieterich (1987). Then
in 1990 four patients from three continents (Australia,
Europe, and North America) were described with
tonic contraversive OTR due to unilateral paramedian mesodiencephalic lesions (Halmagyi et al. 1990).
It was proposed that the tonic contraversive OTR in
these patients was due to persistently decreased resting activity of ipsilateral interstitial nucleus neurons,
in contrast to the paroxysmal ipsiversive OTR
ascribed earlier to transiently increased activity of
the same interstitial nucleus neurons. Only then was
it clear that OTR is not at all a rare condition, but a
frequent syndrome in unilateral brainstem lesions
such as Wallenberg's syndrome, which manifests in
one-third of these patients as a complete OTR (p.
309; Dieterich and Brandt 1992).
Mechanism of OTR

OTR obviously represents a fundamental pattern of


coordinated eye-head roll motion based on utricular
and vertical canal input (Brandt and Dieterich 1987).
In acute lateral medullary lesions it is clearly caused
by the interruption of the ascending otolith and posterior canal inputs ne ar the vestibular nuclei
(Dieterich and Brandt 1992). There is convincing
evidence that OTR occurs with even more peripheral
(labyrinthine and eighth nerve) lesions (Halmagyi et
al. 1979; Dieterich et al. 1989; Arbusow et al. 1997).
Brain (1926) first suspected that OTR or at least a
combination of ipsilateral head tilt with skew deviation in a patient suffering from chronic nonsuppurative otitis media was a syndrome of
peripherallabyrinthine dysfunction.

Fig. 10.4. Tonic ocular tilt reaction in three patients suffering from chronic midbrain lesions. Note sustained head tilt and concurrent
vertical divergence of the eyes (skew deviation). (From Brandt and Dieterich 1987.)

181

Vestibular disorders in (frontal) roll plane

The specific role of the utricle in the generation of


aTR is supported by earlier animal experiments
(Suzuki et al. 1969; Fluur and Mellstrm 1970a,b). By
electrically stimulating single utricular nerves in the
cat, it was possible to induce cyclorotation concurrently with vertical divergence and slight contralateral horizontal shifts of both eyes (spinal cord cut
between Cl and C2 levels). Electrical stimulation
synchronised action potentials in contrast to natural
stimulation, which desynchronises them. The experimentally induced reaction thus showed contraversive ocular torsion as weIl as skew deviation caused
by ipsilateral hypertropia and contralateral
hypotropia. Accordingly, muscle tension increases.
EMG responses of extraocular muscles were
strongest in the ipsilateral superior oblique and contralateral inferior oblique. Neurophysiologically, the
utricle projects to the lateral vestibular nucleus and,
to a minor extent, to the superior vestibular nucleus.
The synaptic organisation of utricular input must
provide patterns of activation for both spinal motor
neurons (head tilt) and conjugate cyclodeviation
with disconjugate vertical divergence. Descending
vestibulospinal projections from the lateral and
medial vestibular nuclei and also descending projections in the MLF to the spinal cord are weIl known.
Several ascending pathways from the vestibular
nuclei complex to the ocular motor nuclei have been
confirmed by lesions and degeneration patterns as
weIl as by the use ofaxoplasmic tracers (Gacek
1971,1982; Lang et al. 1979; Carpenter and Cowie
1985). An ipsilateral projection from the superior
vestibular nucleus follows the MLF and terminates
in the ipsilateral fourth nerve nucleus, the bilateral
oculomotor complex (inferior rectus subdivision),
the ipsilateral nucleus of Darkschewitsch (probably
involved in not only eye movements), the ipsilateral
INC, and the rostral interstitial nucleus of the MLF
(riMLF). Another ipsilateral projection originates
from the lateral vestibular nucleus and terminates
within the ipsilateral oculomotor complex via the
ascending tract of Deiters. Finally, a contralateral
projection, primarily from the medial vestibular
nucleus, ascends in the MLF to terminate in the
third, fourth, sixth, and Darkschewitsch nuclei as
weIl as in the INC. In order to generate the typical
eye movements of the aTR, it could be argued that
particular activation of the ipsilateral inferior rectus
and inferior oblique (ipsilateral ascending tract of
Deiters or via the ipsilateral MLF?) as weIl as the
contralateral superior rectus and superior oblique
(contralateral ascending tract within MLF) are
required. However, contrary to the findings of Gacek
(1971), there is experimental evidence that the
ascending tract of Deiters is not involved in vertical
eye movements but conveys information for the

medial rectus sub division (Reisine and Highstein


1979; Lang et al. 1979; Carpenter and Carleton 1983).
The INC has monosynaptic excitatory and
inhibitory connections to the fourth nerve motor
neurons, a synaptic organisation required of an integration centre for reciprocal control of vertical and
rotational eye movements (Schwindt et al. 1974). It
can coordinate eye and head tilt in roll and pitch by
virtue of its efferent monosynaptic projections, not
only to the ocular but also to spinal motor neurons
(King et al. 1980, 1981). The INC is the velocity-toposition neural integrator for vertical and torsional
eye movements (Fukushima 1991; Crawford et al.
1991).
Second-order type I vestibulospinal neurons are
unlikely candidates for the neuronal mechanisms
subserving (voluntary) directional synergy of eye
and head roll motion, because of the crossed
vestibulocollic reflex connection. Berthoz and
Grantyn (1986) described ipsiversive coupling of eye
and head, which could be mediated by descending
excitatory ipsilaterally projecting tectoreticulospinal
neurons. These authors showed that the complex
axonal branching as weIl as the intense bursts associated with contraversive saccades and contralateral
neck EMG during visual orientation of the animal
are compatible with eye-head synkinesis in roll.
OTR and perceived tilt

The significant tilt of the subjective visual vertical


toward the head tilt in all patients with aTR suggests
that this is the perceptual correlate of an apparent
body tilt relative to the true earth coordinates. There
must be an inclination of the internal representation
of the gravitational vector which causes both the
seen environment and posture to appear tilted contraversive to the lesioned side (Brandt and Dieterich
1987). Both perception and postural control are
affected, and the consequence is a compensatory
ipsiversive eye-head tilt, the aTR (Fig. 10.5). Despite
the resulting postural imbalance and the visual reference of objective vertical, the eyes, head, and body
are adjusted to what the CNS erroneously computes
to be vertical.
Two types of OTR: the medullary "ascending"
VOR-OTR and the mesencephalic "descending"
integrator-OTR

There is sufficient clinical evidence now available to


differentiate between two brainstem types of aTR
(Brandt and Dietrich 1998): the "ascending"
medullary and the "descending" mesencephalic
types. aTR secondary to pontomedullary vestibular
nuclei lesions (i.e. Wallenberg's syndrome) constitutes

182

Vertigo

PERCEIVED TILT

thalamus J ves t ibu [ar cortex


~

"so -J

HYPERTROPIA

E XCYCLOTROPIA

SR

(compensatory)

t~
.......

ocular tilt reaction .......1

:.'.; /

Ci)

CD
normal upright
Fig.l0.5. Ocular tilt reaction (OTR) represented as a "motor
compensation" of a lesion-induced apparent eye-head tilt
(dashed line) which would be opposite in direction to the apparent tilt. Eyes and head are continuously adjusted to what the
lesioned brain computes as being vertical. (From Brandt and
Dieterich 1987.)

a tone imbalance of the VOR in roll, whereas OTR in


INe lesions secondary to paramedian infarctions
may constitute a tone imbalance of the neural integration centre for vertical and rotatory eye-head
co ordination (Dieterich and Brandt 1993b). The different manifestations of the "ascending" VOR type
with monocular or dysconjugate ocular torsion of
the eyes (Figs. 10.6 and 10.7) depend on whether
input from fibres of the posterior, anterior, or both
semicircular canals is involved.
Thus, the "ascending" type ofVOR-OTR with pontomedullary lesions of the medial or superior
vestibular nucleus or graviceptive pathways, wh ich
subserve the VOR in the roll plane, reflects a tone
imbalance of the VOR in roll. If the crossing pontomesencephalic "graviceptive" pathways are
lesioned on one side - rostral to the downwardbranching of vestibulospinal connections - skew torsion results without head tilt (see p. 185).
The "descending" mesencephalic type of integrator-OTR results from lesions of the rostral midbrain
integration centre for eye-head co ordination in roll
and pitch planes. This centre not only integrates eye
and head velo city to position (for holding eye-head
positions in space), but it also coordinates vestibular
reflex responses (VOR) with voluntary cortical control during active locomotion. While a unilateral
mesodiencephalic lesion that inactivates both the
INe and the adjacent riMLF will produce a contralesional tonic OTR (Halmagyi et al. 1990), a unilateral mesodiencephalic lesion that inactivates only the
INe but spares the riMLF will produce a jerk see-saw

apparent tllt

+ ' 10

:~~
~ ..

HC

PC
?

otolithic input

"
~

11 R

"-.

I
I
I
I
I

MLF

I
I
I
I

,
cervical cord
HEADTILT

Fig. 10.6. Hypothetical explanation of ocular tilt reaction due


to lesions of the vertical semicircular canal pathways as demonstrated by a schematic drawing of the three-neuron vestibuloocular reflex arc between the posterior left and the anterior
(fight) semicircular canals and the ocular muscles. An excitatory
ascending pathway is linked from the posterior semicircular
canal to the ipsilateral superior oblique and the contra lateral
inferior rectus muscle (Ieft); an inhibitory ascending pathway is
linked to the ipsilateral inferior oblique and the contra lateral
superior rectus muscle (Graf et al. 1983; Graf and Ezure 1986). A
lesion of these pathways causes excyclotropia of the ipsilateral
and hypertropia of the contralateral eye. An excitatory ascending
pathway projects from the anterior semicircular canal to the ipsilateral superior rectus and the contra lateral inferior oblique
muscle (fight); an inhibitory pathway projects to the ipsilateral
inferior rectus and the contralateral superior oblique muscle. A
lesion of these pathways of the anterior semicircular canal will
cause a hypotropia of the ipsilateral and incyclotropia of the contralateral eye. A combination of both, anterior and posterior
semicircular canal pathways, induces a complete ocular tilt reaction (compare Figure 10.7). AC, PC, and HC = anterior, posterior
and horizontal semicircular canal; MLF = medial longitudinal
fasciculus; SO and 10 = superior oblique and inferior oblique
muscles; SR and IR = superior and inferior rectus muscles (From
Dieterich and Brandt 1992.)

183

Vestibular disorders in (frontal) roll plane

!~~

AC

~~t

~~~t
OS RS

PC
AC/PC

through the MLF (Fukushima et al. 1987) to the pontine level, and couple eye and head roll motion by
excitatory ipsilateral projections with complex axonal
branching (Dieterich and Brandt 1993b). This view is
supported by the binocular ocular torsion which
appears in bilateral chronically labyrinthectomised
cats after unilateral INC deactivation by muscimol
infusion (Fukushima 1991; Fukushima et al. 1992).
We propose the following differentiation between
the two types of OTR (Brandt and Dieterich 1998;
Fig. 10.8):
the "ascending" pontomedullary VOR-OTR with
ipsilateral lesions of the VOR pathways in roll,
dose to the vestibular nudei. This type is characterised by dysconjugate ocular torsion and
occurs if anterior, posterior, or both semicircular
canal and otolithic pathways are affected;
2. the "descending" mesencephalic integrator-OTR
with contralaterallesions of the rostral midbrain
integration centres (INC, riMLF) for eye-head
co ordination in the roll and pitch planes. This
type is characterised by conjugate ocular torsion.
1.

01 RI

111

IV
MLF

HC

Fig. 10.7. Hypothetical explanation of ocular tilt reaction (OTR)


due to lesions of one or two vertical semicircular canal pathways.
Although imbalance may be primarily otolithic, these inputs converge with those of the semicircular canals in the vestibular
nucleus. A schematic drawing of the three-neuron vestibulo-ocular reflex arc between the posterior (PC) and the anterior (AC)
semicircular canals and the extraocular eye muscles. An excitatory ascending pathway is linked from the posterior semicircular
canal to the ipsilateral superior oblique (OS) and the contra lateral
inferior rectus muscle (RI).A lesion ofthis pathway causes excyclotropia ofthe ipsilateral eye and hypertropia ofthe contralateral eye (PC type). An excitatory ascending pathway projects from
the anterior semicircular and the ipsilateral superior rectus (RS)
and the contralateral inferior oblique (01) muscles. Alesion of the
pathway of the anterior semicircular canal would cause a
hypotropia of the ipsilateral and an incyclotropia of the contralateral eye (AC type). A combination of both anterior and posterior semicircular canal pathways induces a complete OTR
consisting of hypotropia of the ipsilateral eye, hypertropia of the
contralateral eye, and conjugate ocular torsion of both eyes in
the direction of head tilt.1I1 = oculomotor nucleus; IV = trochlear
nucleus;VIII = vestibular nucleus; MLF = medial longitudinal fasciculus. (From Brandt and Dieterich 1994.)

nystagmus (Halmagyi et al. 1994; p. 193). The mesencephalic type of OTR is characterised by conjugate
ocular torsion and involves descending pathways,
such as tectoreticulospinal neurons (Berthoz and
Grantyn 1986), which originate in the INC, run

Both effects can be explained by lesional-vestibular


tone imbalance caused by decreased resting activity
of the graviceptive pathways or INC neurons. In contrast, paroxysmal effects of certain structures have
been attributed to transiently increased activity or
electrical stimulation (Rabinovitch et al. 1977;
Hedges and Hoyt 1982; Lueck et al. 1991). The effects
caused by paroxysmal activation are opposite in
direction to those caused by lesional inhibition, such
as unilateral infarction (Halmagyi et al. 1990).
Thus, the current functional concept is that there
are two distinct and separate brainstem structures
controlling eye-head co ordination in roll and pitch
planes:

the bilateral caudal VOR with inputs from the


otoliths and the vertical semicircular canal, and
the bilateral rostral integration centre (INC,
riLMF).

These structures have specialised functions as


regards their dynamic and static (velocity to position) aspects. However, they are not functionally separate but rather they cooperate during active
locomotion and orientation, which require both
rapid responses with reflexive change in position
and maintenance of the achieved position by integration. Consequently, disorders of either structure
will affect both dynamic and static functional
aspects, making it all the more difficult to understand the pathophysiology and dinical dassification.
Contrary to a common belief, head tilt produced

184

Vertigo
eortex

thalamus
" descendlng
Integrator - OTR "

INC, riMlF

; '--....
\ :.....-....,~_

0 ...;.: .., .

-'1
. - - -...,

ocular motor nuciel

skew-torsion
....

vestibular nuelei

" aseending
VOR - OTR "

~~~

utricle and
vertieal semielreular eanals

Fig. 10.8. Schematic presentation of the clinically relevant two different types of ocular tilt reaction (OTR): the "ascending"VOR-OTR
and the "descending" integrator-OTR. The VOR-OTR is induced by lesions or pathological excitation of the otoliths, the vestibular nerve,
or the vestibular nuclei (lateral medulla). Integrator-OTR is induced by rostral midbrain lesions involving the interstitial nucleus of (ajal
and the rostral interstitial nucleus of the medial longitudinal fascicle (lNC, riLMF). Lesions of "ascending" gravitational pathways from
the vertical semicircular canals and the otoliths at pontomesencephalic level (rostral to downward branching of vestibulospinal pathways) cause skew torsion without head tilt. The VOR and the neural integrator in the mesencephalon, however, are two components of
a functionally cooperative system mediating eye-head coordination in pitch and rol l. The VOR is specialised on the dynamic reflexive
response, whereas the integrator maintains position.

by hemilabyrinthectomy in the cat was not dependent on the activity of the direct vestibulospinal
tracts (Fukushima et al. 1988). Hemilabyrinthectomy
produced a characteristic head tilt even in those cats
in which the medial and/or one lateral vestibulospinal tract(s) had been interrupted. Lesions of the
medial vestibulospinal tract did not influence a preexisting head tilt produced by hemilabyrinthectomy.
These results suggested that the head tilt produced
by hemilabyrinthectomy does not depend on the
activity of direct vestibulospinal fibres from the VOR
(Fukushima et al. 1988). One could speculate that the
vestibular tone imbalance due to hemilabyrinthectomy
causes the contralateral INC to produce tonic head
tilt, which is then slowly and cent rally compensated
by other structures. Thus the causative lesion is in
the caudal VOR, but head tilt as a sign is generated
by the rostral integration centre.
It is clinically useful to separate the VOR~OTR
from the integrator-OTR simply because of the different sites and sides of the lesions. However, as
regards the clinical syndrome, it is highly probable
that most of the static deviations in roll are due to
tonic asymmetry of the integrator system, even in
the VOR type of OTR. Functional interaction of the
INC and the riLMF as well as the complex modularity
and parallel processing of the neural integrator
system for eye-head co ordination in vertical planes
(Crawford et al. 1991; Crawford and Vilis 1993) rep-

resent basic sensorimotor control mechanisms. They


are also valid for limb movement control, e.g. interaction between active locomotion and spinal stretch
reflexes.
Aetiology

The two most common causes of tonic OTR are


brainstem ischaemia (especially Wallenberg's syndrome and unilateral paramedian thalamic infarctions) and brainstem tumours (Brandt and Dieterich
1987; Halmagyi et al. 1990). We have also seen cases
with unilateral thalamic haemorrhages, lower brainstern haemorrhages (cavernous angioma, lymphomas), after severe brainstem concussion, in
multiple sclerosis, or associated with attacks of
basilar artery migraine.
The paroxysmal OTR described in a patient with
MS (Rabinovitch et al. 1977) may be a variant of the
paroxysmal attacks assumed to arise from ephaptic
conduction between adjacent demyelinated axons
(see paroxysmal dysarthria and ataxia in MS, p. 242).
This view is supported by the efficacy of carbamazepine treatment of this patient. The patient
with a brainstem abscess described by Hedges and
Hoyt (1982) had paroxysmal attacks strongly reminiscent of the paroxysmal brainstem attacks
described above. Both attacks were characterised by
periodic paroxysms, occurring several times an

185

Vestibular disorders in (frontal) roll plane

hour and lasting 3-5 minutes, and baclofen had


some therapeutic effect. We have observed repeated
paroxysmal attacks of contraversive OTR with
ipsiversive torsional nystagmus in the acute stage of
a case of Wallenberg's syndrome.
Otolithic dysfunction can cause transient tonic
OTR (e.g. following stapedectomy; Halmagyi et al.
1979) or paroxysmal OTR (e.g. otolith Tullio
phenomenon, p. 108; Brandt et al. 1988; Dieterich et
al. 1989). Herpes zoster oticus can cause transient
OTR if it involves the superior and inferior divisions
of the vestibular nerve (Arbusow et al. 1997).
Drug-induced cases of OTR or OTR due to epileptic discharges have not yet been described, but basically seem possible from our understanding of the
organisation of the vestibular system.

IOcular

Torsion

___ RE

20

-fl- LE

-+-- SO
10

o+0--,-----,--,-----,-,-----T==-.----;.1--.----I~8
10

15

20

25

25

30

35

SVV - Tilt

40

45
days

months

___ RE

20

-8- LE

Natural course and management


The natural course and management of OTR depend
on the aetiology. OTR can be transient; in cases of
haemorrhage or infarction recovery occurs within a
few weeks. However, it can be permanent, as we
observed in a patient with severe brainstem concussion. Following mesodiencephalic infarctions, all
features of OTR - postural, ocular motor, and perceptual - disappear naturally and gradually within 5
weeks to months (repeated measurements made in
seven patients = up to 1-3 years; Dieterich and
Brandt 1993b) (Fig.10.9). This is probablybased on a
functionally significant central compensation of a
vestibular tone imbalance induced by a unilateral
centrallesion. The mechanisms underlying central
compensation of centrallesions may be similar to
those of central compensation of peripheral vestibular lesions (Chap. 3; p. 55). Physical therapy may
facilitate this central compensation.
Paroxysmal OTR in MS was treated effectively
with carbamazepine (Rabinovitch et al. 1977);
baclofen had some therapeutic benefit in a patient
with paroxysmal OTR and brainstem abscess
(Hedges and Hoyt 1982). Paroxysmal OTR due to
otolithic Tullio phenomenon (p. 107) was successfully
treated by surgical fixation of the luxated stapes
footplate (Brandt et al. 1988; Dieterich et al. 1989).
Table 10.2 summarises the information given in
this chapter about OTR.

Skew deviation (skew-torsion sign)


Skew deviation of the eyes (or the HertwigMagendie sign) is a common supranuclear vertical
misalignment of the visual axes which is associated
with lesions in the posterior fossa, particularly those
involving the brainstem tegmentum (Smith et al.

10

15

20

25

30

35

40

45
days

8
months

Fig. 10.9. Patient with a left para median infarction presenting


with a complete ocular tilt reaction (OTR) to the right. OTR consisted of contraversive head tilt of 20 (bottom); skew deviation of
10, left eye over right eye; and ocular torsion of 15-20 (counterclockwise from the viewpoint of the observer). Natural course of
ocular torsion, skew deviation (5D), and tilt of subjective visual
vertical (5W, in degrees) shows gradual recovery in 6 weeks. RE =
right eye; LE = left eye. (From Dieterich and Brandt 1993.)

1964; Keane 1975). The numerous studies of mesencephalic, pontine, and medullary lesions have generally concluded that skew deviation is an imprecise
localising sign (Keane 1975; Goldstein and Cogan
1961). Clinical studies using autopsied material are
controversial, for they have demonstrated lesions
ipsilateral (Smith et al. 1964; Keane 1975; Martinez
and Fox 1973) or contralateral (Martinez and Fox
1973) to the lowermost eye at the pontine level. In
contrast, lesions ipsilateral to the lowermost eye
have been consistently reported in patients with
medullary infarctions (Silfverskild 1965; Hagstrm
et al. 1969; Morrow and Sharpe 1988; Dieterich and
Brandt 1992), whereas contralateral lesions have
been described in patients with midbrain dis orders
(Wakusawa 1973; Halmagyi et al. 1990).
The most appealing hypothetical mechanism
explaining skew deviation is that it results from unilateral damage of the tonic otolith-ocular pathways
(Keane 1975) or the (combined otolithic and vertical
canal) "graviceptive" pathways mediating the
vestibulo-ocular reflex (VOR) in the roll plane
(Brandt and Dieterich 1991; Dieterich and Brandt
1992).

Vertigo

186

Table 10.2. Ocular tilt reaction (OTR)


Clinical syndrome
Tonic or paroxysmal eye-head synkinesis consisting of
Lateral head tilt (roll plane)
Deviation of subjective vertical (toward head tilt)
Skew deviation (hypotropia of the undermost eye)
Ocular torsion (toward head tilt)
Two types
Type I
ipsiversive "ascending" VOR-OTR with peripheral vestibular
or central-vestibular lesions at medullary level
(dysconjugated ocular torsion)
Type 11
contraversive "descending" integrator-OTR with central
vestibular lesions at mesencephalic level (lNC, riMLF)
(conjugated ocular torsion)
Incidence/sex
Depends on aetiology, affects both sexes equally
Pathomechanism
Type I VOR-OTR:

tone imbalance of the VOR in roll plane due to


"graviceptive" pathway lesions
Type IIlntegrator-OTR: tone imbalance ofthe bilateral mesencephalic
neural integrator for eye-head coordination in
roll plane
The direction of tilt is ipsiversive for lesions induced by VOR-OTR and
contraversive for integrator-OTR. Directions reverse if OTR is induced by
paroxysmal excitation (or stimulation) ofthese structures

Aetiology
Type I:

Type 11:

otolith, vertical semicircular canals, labyrinth and vestibular


nerve disorders, infarctions (Wallenberg's syndrome)
haemorrhages, tumours, MS plaques of lateral medulla
(vestibular nuclei)
acute rostral mesencephalic lesion (infarctions,
haemorrhages, MS plaques, tumours, ete)

Course / prognosis
Depends on aetiology, e.g. slow spontaneous recovery with dorsolateral
medullary or mesodiencephalic infarctions due to central compensation
Rarely persistent in structural mesencephalic lesions
Management
ParoxysmalOTR:
Tonic OTR:

medical treatment with carbamazepine or


baclofen
causal treatment of responsible disorder
physical therapy

Differential diagnosis
Trochlear palsy, symptomatic head tilt without skew deviation and
ocular torsion, skew torsion without head tilt

Skew torsion: a vestibular brainstem sign of


topographie diagnostie value
Of 155 patients with the clinical diagnosis of acute
unilateral brainstem infarction, 56 patients (36%)
exhibited skew deviation (Table 10.3; Brandt and
Dieterich 1993). These 56 patients were examined for
static vestibular function in the roll plane. Ischaemic
lesions were allocated to the level and side of the
brainstem on the basis of the clinical syndrome and

neuroimaging. Three findings were of clinical relevance (Fig. 10.10):

ipsiversive (ipsilateral
eye was undermost) with caudal pontomedullary
lesions and contraversive (contralateral eye was
lowermost) with rostral pontomesencephalic
lesions.
2. All skew deviations were associated with concomitant ocular torsion toward the undermost
eye.
3. All skew deviations were associated with tilts of
perceived visual vertical (tilts of SVV adjustment
toward the undermost eye as defined by the
direction of torsion of the upper pole of the eye).
1. All skew deviations were

Thus, skew deviation, or more correctly, ocular skew


torsion is a sensitive brainstem sign that has localising and lateralising value (Brandt and Dieterich
1993). There is evidence that the ocular skew torsion
sign indicates a vestibular tone imbalance in the roll
plane secondary to "graviceptive" pathway lesions.
The only difference between OTR and skew torsion is
the absence of head tilt (Fig. 10.10). Sometimes skew
deviation is misinterpreted as a non-vestibular ocular
motor disorder. What Wiest et al. (1996) describe as a
rare "crossed vertical gaze paresis" (monocular elevation paresis and contralateral downgaze paresis)
caused by mesodiencephalic infarction was more
likely a contraversive skew deviation.

Different types of skew deviation


Although all manifest skew deviations seem the
same to the clinician, skew deviation can result from
different combinations of dysconjugate vertical ocular deviations. Evidence has been presented for at
least three different types of skew deviation when it
occurs as a feature of OTR (Fig. 10.11). The otolithic
type (utricle) is characterised by upward deviation
of both eyes with different amplitudes, as described
for otolith Tullio phenomenon in humans (p. 107;
Dieterich et al. 1989). The dorsolateral medullary
type (Wallenberg's syndrome; p. 309) is distinguished by the fact that hypertropia of one eye may
occur while the other remains in the primary position, depending on whether vertical canal input is
affected only by posterior or by both posterior and
anterior semicircular canals (p. 183; Dieterich and
Brandt 1992). The rostral midbrain type simultaneously involves hypertropia of one eye and
hypotropia of the other eye, as described during
electrical stimulation of the midbrain tegmentum in
monkeys (Westheimer and Blair 1975a,b) and
observed in clinical cases with paroxysmal OTR
(Rabinovitch et al. 1977; Hedges and Hoyt 1982).

187

Vestibular disorders in (frontal) roll plane

Table 10.3. Differential effeets of aeute unilateral isehaemie brainstem lesions on skew deviation, oeular torsion (OT), and internuclear ophthalmoplegia (lNO)'
Lesion

Skew

OTin skew

(%)

Binoeular

INO in skew
Lowerrnost eye

Mesodieneephalie

35

(23)

Meseneephalie
Pontomeseneephalie
Pontine
Pontomedullary
Lateral medullary
Total

22
17
46

(36)
(35)
(30)
(23)
(45)
(36)

4
3

3
2
12
20

17
53
155

6
14
4
24
56

1
10
26

INO

Uppermost eye

1
3
3
1
2
10

2e
1e
7(5e,2i)
1i

11

2
6
10
1
19

'Patients with mesodieneephalie lesions (top) do not add to the total at the bottom.
bAdditional third nerve palsy.
N = number of patients with brainstem infaretions; n = number of patients with skew; OT = oeular torsion; lowermost = monoeular OT ofthe lowermost
eye; uppermost = monoeular OT of the uppermost eye; e = contraversive; i = ipsiversive.
From Brandt and Dieterieh (1993).

Pontomesencephalic
(Iesion left)

Pontomedullary
(Iesion left)
Fig. 10.10. Ocular skew-torsion sign in unilateral brainstem
lesions. Skew deviation is always ipsiversive with caudal (pontomedullary) lesions and contraversive with rostral (pontomesencephalic) lesions. Skew deviation is as a rule associated with
ocular torsion (OT) and tilt of perceived vertical toward the
undermost eye.ln medullary lesions, OT is mostly disconjugate
with predominant excyclotropia of the undermost eye, whereas
OT in pontomesencephalic lesions is mostly conjugate. Thus,
skew torsion is a sensitive brainstem sign with localising and lateralising value. (From Brandt and Dieterich 1993.)

Alternating skew deviation

About 12% of the patients with skew deviation


exhibit either hypertropia on lateral gaze, which
alternates to either side, or a natural, slowly alternating skew deviation, which resembles both see-saw
nystagmus and OTR (Corbett et al. 1981; Mitchell et
al. 1981; Keane 1985). See-saw nystagmus has been
reported with lesions of the mesodiencephalic junction (Daroff 1965) or as a congenital abnormality
(Schmidt and Kommerell1974). According to Keane
(1985), one can assume that alternating skew deviation indicates midline or bilaterallesions with pretectal preference. This would explain the missing
INO in these patients, whereas a combination of
skew deviation and INO is frequently seen with unilateral pontine and pontomesencephalic lesions
(Table 10.3).

Mesencephalon

,--..,..

Medulla

Otoliths

Fig. 10.11. Different types of skew deviation have been


described as a feature of ocular tilt reaction (OTR), depending on
the site from which OTR is elicited. Simultaneous hypertropia and
hypotropia have been described for paroxysmal OTR in stimulations of the midbrain tegmentum in the monkey and humans;
monocular hypertropia causes skew deviation in dorsolateral
medullary lesions in humans; concurrent ocular deviation of
both eyes with different amplitudes was observed with mechanical stimulation of the otoliths in humans. Additional torsional eye
movements are indicated by curved arrows; site of stimulation or
lesion is left. (From Brandt and Dieterich 1991.)

Moster and co-workers (1988) proposed another


interpretation of the mechanism of alternating
skew. They found that most patients with this syndrome had associated downbeat nystagmus and
had been diagnosed to have lesions of the cervicomedullary junction or cerebellar pathways. They
believed that lesions of the cerebellum or their projections can cause alternating skew deviation without involving the primary otolithic pathways
(Moster et al. 1988). Caution, however, should be
exercised here because their arguments are based
on animal studies in monkeys (Burde et al. 1975)
that underwent total cerebellectomy and thus also

Vertigo

188

had damaged brainstems. Burde and co-workers


(1975) described in their article that "many of these
animals demonstrated ocular movement dis orders
associated with induced brainstem dysfunction"
and that in serial section examination "many animals were also found to have some damage,
although minimal, to the vestibular nuclei on one
side." Earlier reports on skew deviation in humans
with cerebellar haemorrhages (Brennan and
Bergland 1977; Ott et al. 1974) obviously also
involved both the cerebellum and the brainstem,
but there is clinical evidence available that alesion
restricted to the cerebellum can cause skew
(Mossman and Halmagyi 1997). Zee (1996), however, presented new, convincing theoretical considerations on the mechanism of alternating skew
deviation in patients with cerebellar lesions. He
proposed a mechanism for alternating skews related to the otolithic-ocular response to fore-aft pitch
of the head in lateral-eye animals. He then attributed the emergence of skew deviation in frontal-eye
animals and humans in pathological conditions to

tion that normally corrects for the differences in


pulling directions and strengths of the various
ocular muscles as the eyes change positions in
the orbit.
Such a compensatory mechanism is necessary to
ensure optimal binocular visual function during and
after head motion (Zee 1996). Ihis compensatory
mechanism may depend upon the cerebellum.
It is unlikely that skews secondary to medullary,
upper pontine, mesencephalic, or cerebellar lesions
are produced by identical mechanisms, since evidence
for different types and mechanisms of skew deviation
has been presented (Brandt and Dieterich 1991).
Natural course

Repeated measurements of skew deviation, ocular


torsion (OI), and tilts of perceived vertical (SVV)
made during a single day showed consistent tilts
(Dieterich and Brandt 1993a). Repeated measurements on subsequent days showed a gradual recovery, mostly within 30 days, both for OI and SVV
(Figs. 10.12; 10.13). Some patients, however, maintained a residual OI of a few degrees without a
corresponding tilt of SVV for up to 2 years.

1. an imbalance in otolith -ocular pathways, and


2. a loss of the component of ocular motor innerva-

RIGHT EYE

300

IIIIGMT EllE

200

2&"

11'

zoo
11"

100

1O"

I'

J/\

'Q

S"

10

11

20

21d
5'

I'
100

' \

IVV
.OT

'.

'

..

0-

YD

~~~
.~."'~
... H~/'
~Dr0 ~
~ 0

)J

100

--"

. 0_

0 0'

15'

zoo

11'

zs'
20"

LI" EVI

30'

LEFT EYE

Fig. 10.12. Two representative time courses of deviations of subjective visual vertical (SW) and ocular torsion (OT) (separate for the
left and right eyes) in a patient with Wallenberg's syndrome on the left a and a patient with unilaterallesion of the region of the interstitial nucleus of Cajal (INC) in the rostral midbrain tegmentum b. Note the dissociated effects in the patient with Wallenberg's syndrome; OT and SVV deviated most in the ipsilateralleft eye. Comparison of individual OT and SVV values in both patients shows
varying dissociations of the net tilt. Both tend to normalise within 4-6 weeks, and fluctuations cannot be simply explained by methodological inaccuracy. VD = vertical divergence as skew deviation; d = days; m = months. (From Dieterich and Brandt 1993a.)

Vestibular disorders in (frontal) roll plane


INe LESION LEFT

189

WALLENBERG'S SYNDROME RIGHT


8th day

4th day

@-r:;
30" ex

rigIll.ya

12" In

..tt ey.

rtght eye

121h day

@o
20" ex

!ett eye

121h day

Historical reports on SVV ti/ts

7- in

21.t day

218t day

~o
!5-

.)1

clockwise optokinetically induced ro11vection is usually greater than the tilt of static SVV. The perception
of body verticality (subjective postural vertical)
when sitting on a two-axis rotatory chair seems less
sensitive in peripheral and central vestibular disorders (Bisdorff et a1. 1996). The perception of body
verticality while seated depended mainly on proprioceptive/contact cues, which were susceptible to tiltmediated adaptation.

5- in

Fig. 10.13. Schematic drawings from original fundus photographs made with the head upright, in a patient with a mesodiencephalic infarction on the left involving the region of the
interstitial nucleus of Cajal (INC) (feft) and a lateral medullary infarction on the right (right) in the acute stage (top) and 12 days and 21
days later, respectively (bottom). The anatomical structures
involved were identified with MRI projections onto cytoarchitectonic sections of a stereotaxic brainstem atlas. Both binocular
pathological ocular torsion (OT) in the mesodiencephalic lesion
and the predominantly ipsilateral excyclotropia (ex) in the lateral
medullary lesion exhibit spontaneous recovery to physiological
excyclotropia of about 4-7 degrees.After 21 days the patient with
Wallenberg's syndrome showed a nearly normalised eye position
(excyclotropia of 6 of the left eye and of 9 of the right eye), while
the patient with the INC lesion on the left had a slight incyclotropia
(in) of the left eye of 5. (From Dieterich and Brandt 1993a.)

Perceived vertical (subjective visual vertical)


In our study 94% of patients with acute unilateral
brainstem infarctions exhibited pathological tilts of
the static subjective visual vertical (SVV) from the
true vertical (Table 10.1; Fig. 10.2; Dieterich and
Brandt 1993a). The mean net tilt angle was about 8
(range of 2-26). It is striking that caudal brainstem
lesions cause ipsiversive tilts of SVV, whereas upper
brainstem lesions cause contraversive tilts. The
direction of SVV tilt corresponds in a11 cases to the
direction of ocular torsion (Fig. 10.14), if present.
Deviations are greatest in cases of lateral medu11ary
lesions (Wa11enberg's syndrome) and mesencephalic
lesions affecting the INC. Adjustments of SVV under
dynamic conditions (ro11vection) confirm the tilt
found under static conditions. The resulting asymmetry of dynamic SVV under clockwise or counter-

The older literature on subjective visual vertical and


visual horizontal with peripheral vestibular, brainstern, and cortical lesions presents contradictory
results. None of these studies focused on vestibular
function in the roll plane, and none provided sound
neural anatomical correlations. Friedmann (1970),
who convincingly described ipsiversive deviations of
the SVV with peripheral vestibular and brainstem
lesions, did not find significant deviations in patients
with corticallesions. Bender and Jung (1948) found
contraversive adjustments of SVV in World War II
veterans with brain injuries in which the frontal or
parietal regions were affected, whereas Teuber and
Mishkin (1954) failed to confirm these findings in
patients studied up to 10 years after injury was
incurred. These and other authors reported greater
variability of SVV adjustments and inconsistent
directional spatial misperception in patients with
corticallesions, hemiplegia, and hemi-Parkinson's
disease.
SVV ti/t - a vestibular sign?

A direction-specific tilt of SVV is a typical vestibular


sign. Support for this is found in psychophysical
studies of patients before and after unilateral
vestibular neurectomy (Curthoys et a1. 1991).
Moreover, our clinica1 experience has shown that it is
frequently associated with ocular torsion, skew deviation of the eyes, and head-and-body tilt in the same
direction.
Lateropulsion of the body is a well-known transient feature of lateral medullary infarction (Hrnsten
1974) in which the patients cannot prevent ipsiversive lateral falls. We believe that subjective vertigo is
usua11y absent in these patients, because there is no
sensory mismatch. The lesion causes deviation of the
perceived vertical, an obligatory sign in Wa11enberg's
syndrome. Individual multisensory regulation of
posture is then adjusted, not to the true vertical but
to the pathologica11y deviated internal representation
of the verticality produced by the lesion. Posturographic measurements in 36 patients with
Wa11enberg's syndrome (Dieterich and Brandt 1992)
revealed that the preferred body-sway direction in

190

Vertigo

subjective
visual
vertical t 5.20

subjective
visual
verticaI7.3

objective
vertical

case 1: 8.S.

right

eve

case 3: G.J.

teil eve

Fig.10.14. Ocular torsion of the hvpotropic eye (fundus photographs with head upright) and deviation of the subjective visual vertical (SVV) with monocular vision in case of ocular tilt reaction (OTR) due to chronic midbrain lesions.ln order to make a quantitative
comparison possible between the angles of ocular torsion and the perceived vertical, the adjustments of the SW were projected onto
the fundus. The patients adjusted SW to 15.20 clockwise and to 7.3 0 counterclockwise. The tilt of SW is greater with greater angles of
ocular torsion, but there is no exact quantitative correspondence between the net tilt angle of both.

minor lateropulsion is diagonal; it becomes more lateral in severe cases, particularly when the eyes are
closed (Fig. 19.3). All patients exhibited significant
tilts of the internal representation of the gravity vector, as indicated by deviations of SVV ipsiversive to
the lesion. If one correlates the grade of lateropulsion
with net tilt angles of SVV, it is evident that the more
pronounced the lateropulsion is, the greater is the
spatial disorientation with respect to verticality (Fig.
19.4).

SVV tilt versus room tilt illusion

Transient upside-down inversion of vision (mirror


image; room tilt illusion) has been repeatedly
described in patients with lower brainstem infarction (Hrnsten 1974; Ropper 1983; Slavin and
Lopinto 1987; Steiner et al. 1987; Tiliket et al. 1996)
or with cortical lesions (Smith 1960; Solms et al.
1988), especially in vestibular epilepsy (p. 233). The
brainstem type is caused by infarcts in the territory
of either the medial branch of the posterior inferior
cerebellar artery (Charles et al. 1992) or the anterior
inferior cerebellar artery (Lopez et al. 1995). These
illusions last for seconds, minutes, or up to 2 days.
They are often associated with rotational vertigo at
the beginning, and recovery is either rapid or there
is gradual uprighting to normal position. Case
descriptions suffer from imprecise localisation of
causative structures, but the vestibular nuclei seem
to be a critical structure (Ropper 1983). There is no

agreement about the laterality of complete (visual


environmental) room tilts, which have been
described toward as weIl as away from the side of
vestibular nuclei lesions (Lopez et al. 1995).
SVV tilts and room tilt illusions, such as transient
upside-down vision or 90 tilts, are obviously
vestibular signs that indicate a misperception of verticality. They both involve "graviceptive" pathways
that extend from the otoliths and the semicircular
canals through the vestibular nuclei and the

thalamus to the parietoinsular vestibular cortex


(Chap. 13). Direction-specific SVV tilts have been
reported with lesions of all these structures (Brandt
and Dieterich 1994). Likewise, room tilt illusions
have been reported with either brainstem or cortical
lesions (Tiliket et al. 1996). The question arises
whether both phenomena are the same (Larmande
et al. 1994), their net tilt angles differing only in size.
In our opinion Tiliket et al. (1996) are correct that
room tilt illusion is the manifestation of a dynamic
mismatch between inappropriate otolithic input and
correct visual input. In their study they describe three
patients with skew deviation due to unilateral brainstern lesions, in whom rotatory chair acceleration
about a vertical or an off-vertical axis elicited transient
illusory room tilts of 90 either in the roll or pitch
plane: "the room was seen as being tilted 90 in the
anteroposterior plane and the front wall was perceived
as the ceiling. For another patient, the room was seen
as being tilted 90 toward the left in the frontal plane
and the wall was perceived as the ceiling".

Vestibular disorders in (frontal) roll plane

191

Why are SVV tilts and room tilt illusions different


phenomena? There are a few typical differences
(Brandt 1997):

SVV tilts are usually stable (chronic) signs and


recovery occurs gradually, within days to weeks;
room tilt illusions are paroxysmal or transient
phenomena.

svv tilts manifest as a continuum of angle of tilt

___Jl--1---=F::~:-r---~--~

up to about 30 as a maximum; room tilt illusions occur in 90 steps as a lateral, fore-aft tilt or
upside-down vision.
SVV tilts are not usually associated with the perception of room tilt.

We believe that SVV tilt is a sensitive measure (in


degrees) of a graviceptive vestibular tone imbalance
when determined without additional visual cues of
orientation. SVV tilts are determined under visual
laboratory conditions that do not allow for the visual
spatial cues of horizontal or vertical. A randomly
off-verticalline usually has to be adjusted to subjective vertical in total darkness or in front of unstructured surroundings. Since patients with significant
tilts of the SVV do not usually complain spontaneously about an apparent tilt of the visual scene,
this deviation must be largely compensated by cortical visual-vestibular interaction. The resulting visual
orientation largely reftects this interaction.
Spatial orientation and postural control are based
on multisensory integration, especially visualvestibular interaction (Dichgans and Brandt 1978).
Both senses provide us with cues about vertical orientation in a 3-D co ordinate system. The visual and
vestibular cortices have to match vestibular spatial
coordinates in three dimensions with the orientation
of the visual scene to determine the unique perception of right and left, up and down, and fore and aft
(Brandt 1997). This matching of two different sensory
3-D co ordinate maps is a continuous process occurring under static and dynamic conditions. This
process must be robust to tolerate slight incongruencies between expected and actually received sensory
inputs. Furthermore, it must be plastic in order to
compensate for a chronic tone imbalance (e.g. unilaterallabyrinthine loss) or to adapt to unusual environments (microgravity). The type of plasticity
exhibited by this visual-vestibular interaction has
been best demonstrated by wearing reversing prisms
for weeks (Kohler 1956). This causes not only an
adaptive "uprighting" of the visual scene, but also a
reversal of the VOR (Melvill-Jones 1977).
In brief, room tilt illusions are, in our opinion,
transient mismatches of the visual and vestibular 3-D
map coordinates that occur in 90 or 180 steps as the
erroneous result of the attempted match (Brandt
1997; Fig. 10.15). They are usually induced when the

X ----.

Fig. 10.15. Schematic representation of the head as a cube


with the cortical matching of the vestibular and the visual 3-0
coordinate maps. The three major planes of action of the vestibular system are the frontal roll, the horizontal yaw, and the sagittal
pitch about the x, y and z axes, respectively. (Top) Visual scene
matched with the vestibular coordinates; (middle) room tilt illusion with 180' tilted visual scene in the pitch plane (upside-down
vision); (bottom): room tilt illusion with 90' tilt in the roll plane
(From Brandt 1997).

192
:~~~e~~;~;

Vertigo
Frequency of pathological tilts of subjective visual vertical (SVV) and ocular torsion (OT) in patients with acute unilateral vascular brain-

Brainstem level

OT

SVV
~~------

-~------

Static
-~-

Mesencephalic
Pontomesencephalic
Pontine
Pontomedullary
Lateral medullary
Total

16
12
34
13
36
111c

16
12
34
13
36
111

Dynamic
--

15 (94%)
11 (92%)
31 (91%)
13 (100%)
34(94%)
104(94%)

16 (100%)
12 (100%)
34(100%)
13 (100%)
36 (100%)
111 (100%)

----

13
11
30
10
22
86

Monocular

Binocular

Total
~-

7 (54%)a
7 (64%)a
14(47%)
6(60%)
6(27%)
40(47%)

5 (38%)
2(18%)
10(33%)
2(20%)
12 (55%)b
31 (36%)

-~~-

12(92%)
9(82%)
24(80%)
8(80%)
18 (82%)
71 (83%)

'See text for influence of concurrent trochlear or oculomotor nuclei palsies.


bDysconjugate OT in 6 of 12 patients with binocular OT.
'16 patients were found .to have the first manifestation of probable multiple sclerosis with a single circumscribed plaque in MRI scan
N = total number of patlents; n = number of examinations of SVV/OT.
.
From Dieterich and Brandt (1993a)

vestibular input (preferably dynamic changes) deviates from normal. This may occur as a physiological
response to microgravity (Glasauer and Mittelstaedt
1992) or lesional or epileptic dysfunction in neurological patients. The subsequently perceived tilt in
one plane causes the afflicted person to attribute
upright to horizontal or even down. The visual scene,
how~ver, which itself contains numerous empirical
spatlal cues for upright, will then in turn dominate
and "correct" the perception and spatial orientation.
Vision will tell the vestibular system where upright iso
Consequently, room tilt illusions are transient: you
cannot perceive two verticals at once.

SVV tilts in central vestibular versus peripheral


ocular motor lesions
Misleading combinations of peripheral ocular motor
and ce~tral. vesti~)Ular pathway lesions of SVV may
occur m mldbram tegmentum dis orders involving
the oculomotor or trochlear nuclei fascicles.
Dissociated or monocular deviations of SVV tilt may
then result. Third and fourth cranial nerve palsies
cause only minor and unpredictable monocular SVV
tilts, in contrast to the binocular and conjugate tilts
frequently seen in patients with acute unilateral
brainstem lesions (Dieterich and Brandt 1993c).
Monocular measurements of SVV in patients with
acute bilateral (symmetrical) fourth cranial nerve
palsies showed tilts of similar degrees for both eyes
but in opposite directions. Thus, even abilateral
infranuclear ocular motor dis order with cyclorotation can be differentiated from a supranuclear disorder. The latter is characterised by SVV tilts under
~inocular viewing conditions, and conjugate SVV
hIts, under monocular viewing conditions - both are
tilted in the same direction (Dieterich and Brandt
1993a).

Thalamic and cortical astasia associated with


SVVtilts
An association of SVV tilts with falls is also typical
for posterolateral (vestibular subnuclei) thalamic
lesions. Thalamic astasia (Masdeu and Gorelick
1988) is a condition in which patients without paresis
or sensory or cerebellar deficits are unable to maintain an unsupported, upright posture. Postural
imbalance with a transient tendency to fall has been
reporte.d following therapeutic thalamotomy and
thalamlc haemorrhage (Verma and Maheshwari
1986). According to our experience in some 30
patients with thalamic infarctions, the posterolateral
~ype ~~y cau~e contraversive or ipsiversive postural
mstabdlty (Flg. 19.11) with SVV tilts, whereas the
paramedian type (if it extends into the rostral midbrain) always causes contraversive falls. Astasia and
gait failure with damage of the pontomesencephalic
(locomotor) region was described by Masdeu et al.
(1994). Although not discussed, it could also be
explained in part by a vestibular tone imbalance in
roll, especially since skew deviation was described as
a feature of the syndrome. f 31 patients with infarctions of .the. middle cerebral artery territory, 21
showed slgmficant, mostly contraversive, pathologic~l SVV tilts (Brandt et al. 1994; 1995). The overlappmg area of these infarctions centred on the
posterior insula, which is probably homologous to
the parietoinsular vestibular cortex (Grsser et al.
1990a,b; p. 220). SVV tilts caused by vestibular cortex lesions mayaiso be associated with (a compensatory) body lateropulsion. This explains the cortical
phenomenon "pusher", which physical therapists
readily recognise.

Vestibular disorders in (frontal) roll plane

Torsional nystagmus
The "graviceptive" input from the otoliths converges
with that from the vertical semicircular canals to
subserve static and dynamic vestibular function in
roll. This combination of static and dynamic effects
(Merfeld et al. 1996) is not surprising if one considers
how these functions are corroborated. Our studies
on OTR, lateropulsion, and SVV were concerned
with static effects of vestibular dysfunction in roll
(Brandt and Dieterich 1994,1995). These effects persist for days to weeks, during which time they spontaneously subside. In the acute stage of infarction,
additional dynamic signs and symptoms occur
which consist of lateral rotational vertigo and torsional nystagmus (Morrow and Sharpe 1988; Lopez
et al. 1992). Fast phases of rotational nystagmus are
contraversive in pontomedullary lesions, whereas
the slow phases correspond in direction to the static
deviation.
Several distinct and separate lesions (Figs C.2-3)
have been associated with torsional nystagmus, e.g.
lesions of the vestibular nuclei (Lopez et al. 1992;
Lawden et al. 1995), the lateral medulla (Morrow and
Sharpe 1988; Bttner et al. 1995), in rare cases the
MLF (as indicated by an association with internuclear ophthalmoplegia (Dehaene et al. 1996;
Noseworthy et al. 1988)), the INC, and the riMLF
(HeImchen et a1.1996; Henn 1992; Halmagyi et al.
1994). Fast phases of torsional nystagmus are

contraversive in pontomedullary lesions and


ipsiversive in paramedian pontine and mesencephalic (INC) lesions (rare exception: contraversive in riMLF lesion).

Jerk-waveform see-saw nystagmus (a torsional nystagmus with elevation of the intorting eye and
depression of the extorting eye) is induced by an
inactivation of the INC in the rostral midbrain and is
also ipsiversive (Halmagyi et al. 1994).
The different locations of lesions causing different
directions of torsional nystagmus first appear to be
confusing. They can, however, be explained by the
tonic torsional shift of eye position along the graviceptive pathways from the vestibular nuclei to the
INC. Alesion of the (medial or superior) vestibular
nucleus causes an ipsiversive tonic deviation, i.e.
ipsiversive ocular torsion, with compensatory fast
phases of the torsional nystagmus to the contralesional side. In view of the fact that the pathway
within the MLF crosses to the contralateral side, an
MLF lesion in the pontine and pontomesencephalic
brainstem induces a tonic contraversive deviation
and therefore a torsional nystagmus with the fast
phases ipsilesional. The same is true for alesion of

193

the INC. The only exception to these directional


rules for tone imbalance along the vestibular graviceptive pathways is the riMLF, alesion of which
causes a (non-vestibular?) ocular motor tone imbalance in the opposite direction.
For better understanding, a distinction has to be
made between a vestibular tone imbalance in roll
(typically a medial or superior vestibular nucleus
lesion as in Wallenberg's syndrome) and (nonvestibular) ocular motor imbalance in roll, e.g. secondary to a neural integrator defect (INC) or a
directional deficiency of vertical and torsional saccade generation (riMLF). The latter represents a
purely ocular motor sign without accompanying
vestibular symptoms such as vertigo or postural
imbalance. These two types of torsional nystagmus
correspond to the "ascending" pontomedullary VOROTR and "descending" mesencephalic integratorOTR as described earlier (p. 181). A clear separation
of the vestibular and the ocular motor system, however, is difficult, since some ascending pathways such
as the MLF convey both functions. This is demonstrated clinically in patients presenting with combinations of internuclear ophthalmoplegia, ocular
torsion, and perceptual tilt (Brandt and Dieterich
1993).
The natural course of vestibulo-ocular and "nonvestibular" ocular motor signs may be different. The
vestibular torsional nystagmus in Wallenberg's syndrome, for example, is transient (Morrow and
Sharpe 1988; Dieterich and Brandt 1992), whereas a
torsional nystagmus secondary to rostral midbrain
lesions may be permanent (HeImchen et al. 1996;
Halmagyi et al. 1994). In torsional nystagmus due to
lower brainstem lesions, vestibular compensation,
which takes advantage of the commissural fibres
between the vestibular nuclei, may account for
recovery. This does not occur after a defect of the
"torsional integrator" in rostral midbrain lesions.
A functional separation of the generation of vertical and torsional saccades (riMLF), on the one hand,
and the conversion of this burst into a vertical and
torsional position signal (INC), on the other, have
been elaborated in monkeys (Crawford et al. 1991;
Henn 1992). In riMLF there is lateralisation of function throughout the torsional fast eye movement
generating system, and unilateral inactivation results
in a loss of ipsidirectional torsional fast-phase eye
movements (Riordan-Eva et al. 1996). These findings
allow a functional interpretation of cases in which
either rostral midbrain lesions present with jerk seesaw nystagmus (Halmagyi et al. 1994) or there is an
occasional revers al of torsional nystagmus, depending on which rostral midbrain structure is mainly
involved, the INC or the riMLF (HeImchen et al.
1996). It is important to emphasise that unilateral

Vertigo

194

lesions of the INe cause only contraversive (in case


of excitation, ipsiversive) tilts with ipsilesional torsional nystagmus, whereas a unilateral riMLF lesion
is capable of initiating contralesional torsional
nystagmus.
Furthermore, it was speculated that the signals
used for vertical smooth pursuit are, at some stage,
encoded in a semicircular canal VOR co ordinate
framework (FitzGibbon et al. 1996). There is dose
crosstalk and interaction between the vestibular and
the "non-vestibular" ocular motor system. To illustrate, for the vertical semicircular canals, vertical
and torsion al motion are combined in the same cells,
with the anterior semicircular canals mediating
upward movement and the posterior semicircular
canals mediating downward movements (FitzGibbon
et al. 1996). For the right labyrinth, however, both
vertical semicircular canals produce dockwise slow
phases (ipsilateral eye intorts, contralateral eye
extorts). The opposite is true for the vertical semicircular canals in the left labyrinth; counterdockwise slow phases are produced. Hence, to generate a
pure vertical VOR, the anterior or posterior semicircular canals on both sides of the head must be excited
so that the opposite-directed torsional components
cancel. Thus, if pursuit were organised similarly to
VOR, pure vertical pursuit would require that oppositely directed torsional components cancel in normals. If this did not happen, a residual torsional
nystagmus could appear during attempted vertical
pursuit (FitzGibbon et al. 1996).

Three-dimensional modelling of
static vestibulo-ocular brainstem
syndromes
Static vestibulo-ocular brainstem syndromes are
characterised by skew deviation, a vertical disconjugacy of the eyes, and ocular torsion. These are the
result of a vestibular tone imbalance in the frontal
(roll) plane. The influence of gravity, which is mediated by the utrides, causes similar physiological
changes in static eye position: ocular counterroll,
and conjugate deviations of vertical eye position.
These observations prompted an approach using the
model described he re (Fig. 10.16). On the basis of the
known deviations of static eye position, we devised a
3-D mathematical feedforward model of otolithocular pathway function which also takes into consideration the detailed anatomy of the brainstem
(Glasauer et al. 1998). This model is able to explain
and predict the differential effects of unilateral and

eye position

ATD

BC

gravity
Fig.l0.16. Sketch ofthe 3-D mathematical feedforward model
of otolith-ocular pathway function. Model input is gravity relative
to the head, model output is eye position. Connections are shown
for the left eye only; left utricle (U), left vestibular nucleus (VN),
and their connections are depicted in grey. Further abbreviations:
oculomotor (111), trochlear (IV), abducens (VI) nuclei; mediallongitudinal fasciculus (MLF), ascending tract of Deiters (ATD), brachium conjunctivum (BC); medial rectus (MRl, lateral rectus (LR),
superior rectus (SR), inferior rectus (IR), superior oblique (SO), inferior oblique (10) eye muscles. ln order to weight the "graviceptive"
input in the left oculomotor nucleus, parallel pathways are
assumed to carry equal weights (0.5) and to be summed in the
oculomotor nucleus, wh ich then projects to the eye muscles.
(From Glasauer et al. 1998.)

bilateral, peripheral or central vestibular lesions on


static eye position in roll, pitch, and yaw planes (Fig.
10.17).
The neurological interest in further development
of this model is obvious: it could conceivably allow
us to identify patients with complete or incomplete
vestibular pathway lesions that the model predicts in
various kinds of simulations. Numerous patients
with Wallenberg's syndrome or pontomesencephalic
lesions have been described who presented with different combinations of skew deviation with monocular or binocular torsion (Dieterich and Brandt

Vestibular disorders in (frontal) roll plane

Fig. 10.17. Ocular torsion (top), skew deviation, and horizontal


exotropia (middle) in a 61-year-old patient with acute
Wallenberg's syndrome, which affected the left medial vestibular
nucleus. Comparison of ocular deviations in the patient with
model simulation (bottom). Ocular torsion (top) is shown by fundus photographs of both eyes as the deviation of the papillamacula meridian (diagonal lines) from horizontal. The patient
showed leftward ocular torsion (right eye: incyclotropia of 7 deg;
left eye: excyclotropia of 13 deg), skew deviation right over left (5
deg), and horizontal exotropia (1.5 deg). The left eye also shows a
Horner syndrome.Ocular deviations predicted by the model (bottom) correspond to all clinical deviations: leftward ocular torsion
(right eye: incyclotropia of 1.2 deg; left eye: excyclotropia of 10.4
deg), skew deviation right over left (11.5 deg), and horizontal
exotropia (0.2 deg). (From Glasauer et al. 1998.)

1992; Brandt and Dieterich 1994). However, the attribution of these different ocular motor abnormalities
to the vestibular and ocular motor structures that
are assumed to be affected must still be proven.

References
Angelaki DE, Bush GA, Perachio AA (1993) Two-dimensional
spatio-temporal coding of linear acceleration in vestibular
nuclei neurons. J Neurosci 13:1403-1417
Arbusow V, Dieterich M, Strupp M, Dreher A, Jger L, Brandt Th
(1997) Herpes zoster neuritis involving superior and inferior
parts of the vestibular nerve causes ocular tilt reaction. Neuroophthalmology 19:17-22
Baker R, Precht W, Berthoz A (1973) Synaptic connections to
trochlear motoneurons determined by individual vestibular
nerve branch stimulation in the cat. Brain Res 64:402-406

195
Bartorelli C (1942) Esperienze di stimolazione elettriae della
regione mesencefalica. Arch FisioI42:384-414
Bender M, Jung R (1948) Abweichung der subjektiven optischen
Vertikalen und Horizontalen bei Gesunden und Hirnverletzten.
Arch Psychiatrie 181:193-212
Berthoz A, Grantyn A (1986) Neuronal mechanisms underlying
eye-head coordination. In: Freund HJ, Bttner U, Cohen B, Noth
J (eds) The oculomotor and skeletal motor systems. Progress in
Brain Research Vo164, Elsevier, Amsterdam, pp 325-343
Bisdorff AR, Wolsley CJ, Anastasopoulus D, Bronstein AM, Gresty
MA (1996) The perception of body verticality (subjective postural vertical) in peripheral and central vestibular disorders.
Brain 119:1523-1534
Brain WR (1926) On the rotated or "cerebellar" posture of the
head. Brain 49:61-75
Brandt Th (1997) Cortical matching ofvisual and vestibular 3-D
coordinate maps. Ann NeuroI42:983-984
Brandt Th, Dieterich M (1987) Pathological eye-head coordination in roll: tonic ocular tilt reaction in mesencephalic and
medullary lesions. Brain 110:694-666
Brandt Th, Dieterich M (1991) Different types of skew deviation. J
Neurol Neursurg Psychiatry 54:549-550
Brandt Th, Dieterich M (1993) Skew deviation with ocular torsion:
a vestibular sign of topographie diagnostic value. Ann Neurol
33:528-534
Brandt Th, Dieterich M (1994) Vestibular syndromes in the roll
plane: topographie diagnosis from brainstem to cortex. Ann
NeuroI36:337-347
Brandt Th, Dieterich M (1995) Central vestibular syndromes in
roll, pitch, and yaw planes: topographie diagnosis of brainstem
dis orders. Neuro-ophthalmology 15:291-303
Brandt Th, Dieterich M (1998) Two types of ocular tilt reaction:
the 'ascending' pontomedullary VOR-OTR and the 'descending'
mesencephalic integrator-OTR. Neuro-ophthalmology 19:83-92
Brandt Th, Dieterich M, Fries W (1988) Otolithic Tullio phenomenon typically presents as paroxysmal ocular tilt reaction. Adv
Oto-Rhino-LaryngoI42:153-156
Brandt Th, Dieterich M, Danek A (1994) Vestibular cortex lesions
affect perception of verticality. Ann Neurol 35:528-534
Brandt Th, Btzel K, Yousry T, Dieterich M, Schulze S (1995)
Rotational vertigo in embolie stroke of the vestibular and auditory cortices. Neurology 45:42-44
Brennan RW, Bergland RM (1977) Acute cerebellar hemorrhage.
Analysis of clinical findings and outcome in 12 cases.
Neurology 27:527-532
Burde RM, Stroud MH, Roper-Hall G, Wirth FP, O'Leary JL (1975)
Ocular motor dysfunction in total and hemicerebellectomized
monkeys. Br J OphthalmoI59:560-565
Bttner U, HeImchen CH, Bttner-Ennever JA (1995) The localizing
value of nystagmus in brainstem dis orders. Neuro-ophthalmology
15:283-290
Carpenter MB, Carleton SC (1983) Comparison of vestibular and
abducens internuclear projections to the medial rectus subdivision of the oculomotor nucleus in the monkey. Brain Res
274:144-149
Carpenter MB, Cowie RJ ( 1985) Connections and oculomotor
projections of the superior vestibular nucleus and cell group
"y". Brain Res 336:265-287
Charles N, Froment C, Rode G, Vighetto A, Turjman F, Trillet M,
Aimard G (1992) Vertigo and upside down due to an infarct in
the territory of the medial branch of the posterior inferior cerebellar artery caused by disseetion of a vertebral artery. J Neurol
Neurosurg Psychiatry 55:188-189
Cohen B, Suzuki JI, Bender MB (1964) Eye movements from semicircular canal nerve stimulation in the cat. Ann Otol (St. Louis)
73:153-169
Corbett JJ, Schatz NI, Shults WT, Behrens M, Berry RG (1981)
Slowly alternating skew deviation: description of apretectal
syndrome in three patients. Ann NeuroI1O:540-546

196
Crawford JD, Vilis T (1993) Modularity and parallel processing in
the oculomotor integrator. Exp Brain Res 96:443-456
Crawford JD, Cadera W, Vilis T (1991) Generation of torsion al and
vertical eye position signals by the interstitial nucleus of Cajal.
Science 252:1551-1553
Curthoys IS (1987) Eye movements produced by utricular
and saccular stimulation. Aviat Space Environ Med
58(Suppl):AI92-AI97
Curthoys IS, Dai MI, Halmagyi GM (1991) Human ocular position
before and after unilateral vestibular neurectomy. Exp Brain
Res 85:215-218
Daroff RB (1965) See-saw nystagmus. Neurology 15:874-877
Dehaene I, Casselman JW, D'Hooghe M, Van Zandijcke M (1996)
Unilateral internuclear ophthalmoplegia and ipsiversive torsional nystagmus. J NeuroI243:461-464
Dichgans J, Brandt Th (1978) Visual-vestibular interaction: effects
on self-motion perception and postural control. In: Held R,
Leibowitz HW, Teuber H-L (eds) Handbook of sensory physiology, vol.VIIl. Perception, Springer, Berlin Heidelberg New York,
pp 755-804
Dieterich M, Brandt Th (1992) Wallenberg's syndrome.
Lateropulsion, cyclorotation and subjective visual vertical in
thirty-six patients. Ann NeuroI31:399-408
Dieterich M, Brandt Th (1993a) Ocular torsion and tilt of subjective visual vertical are sensitive brainstem signs. Ann Neurol
33:292-299
Dieterich M, Brandt Th (1993b) Thalamic infarctions: differential
effects on vestibular function in the roll plane (35 patients).
Neurology 43:1732-1740
Dieterich M, Brandt Th (1993c) Ocular torsion and perceived vertical in oculomotor, trochlear and abducens nerve palsies. Brain
116:1095-1104
Dieterich M, Brandt T, Fries W (1989) Otolith function in man.
Results from a case of otolith Tullio phenomenon. Brain
112: 1377-1392
FitzGibbon EJ, Calvert PC, Dieterich M, Brandt Th, Zee DS (1996)
Torsional nystagmus during vertical pursuit. J NeuroOphthalmoI16:79-90
Fluur E, Mellstrm A (1970a) Utricular stimulation and oculomotor
reactions. Laryngoscope 80: 170 1-1712
Fluur E, Mellstrm A (1970b) Saccular stimulation and oculomotor
reactions. Laryngoscope 80:1713-1721
Friedmann G (1970) The judgement ofthe visual vertical and horizontal with peripheral and central vestibular lesions. Brain
93:313-328
Fukushima K (1991) The interstitial nucleus of Cajal in the midbrain reticular formation and vertical eye movement. Neurosci
Res 10:159-187
Fukushima K, Fukushima J, Terashima T (1987) The pathways
responsible for the characteristic head posture produced by
lesions of the interstitial nucleus of Cajal in the cat. Exp Brain
Res 68:88-102
Fukushima K, Fukushima J, Kato M (1988) Head tilt produced by
hemilabyrinthectomy does not depend on the direct vestibulospinal tracts. Brain Behav EvoI32:181-186
Fukushima K, Ohashi T, Fukushima J, Kase M (1992) Ocular torsion produced by unilateral chemical inactivation of the interstitial nucleus of Cajal in chronically labyrinthectomized cats.
Neurosci Res 13:301-305
Gacek RR (1971) Anatomical demonstration of the vestibuloocular projections in the cat. Laryngoscope 81:1559-1595
Gacek RR (1982) The anatomical-physiological basis for vestibular function. In: Honrubia V, Brazier MAB (eds) Nystagmus and
vertigo: Clinical approaches to the patient with dizziness.
Academic Press, New York, London, pp 3-23
Glasauer S, Mittelstaedt H (1992) Determinants of orientation in
microgravity. Acta Astronautica 27: 1-9
Glasauer S, Dieterich M, Brandt Th (1998) Three-dimensional
modeling of static vestibulo-ocular brainstem syndromes.
NeuroReport 9:3841-3845

Vertigo
Goldstein JE, Cogan DG (1961) Lateralizing value of ocular motor
dysmetria and skew deviation. Arch Ophthalmol66:5 17 -518
Graf W, Ezure K (1986) Morphology of vertical canal related second order vestibular neurons in the cat. Exp Brain Res 63:35-48
Graf W, McCrea RA, Baker R (1983) Morphology of posterior
canal-related secondary vestibular neurons in rabbit and cat.
Exp Brain Res 52:125-138
Grsser O-J, Pause M, Schreiter U (1990a) Localization and
responses of neurons in the parieto-insular vestibular cortex of
awake monkeys (Macaca fascicularis). J PhysioI430:537-557
Grsser O-J, Pause M, Schreiter U (1990b) Vestibular neurons in
the parieto-insular cortex of monkeys (Macaca fascicularis):
visual and neck receptor responses. J PhysioI430:559-583
Hagstrm L, Hrnsten G, Silfverskjld BP (1969) Oculostatic and
visual phenomena occurring in association with Wallenberg's
syndrome. Acta Neurol Scand 45:568-582
Halmagyi GM, Gresty MA, Gibson WPR (1979) Ocular tilt reaction with peripheral vestibular lesion. Ann NeuroI6:80-83
Halmagyi GM, Brandt Th, Dieterich M, Curthoys IS, Stark RJ, Hoyt
WF (1990) Tonic contraversive ocular tilt reaction due to unilateral meso-diencephalic lesion. Neurology 40:1503-1509
Halmagyi GM, Aw ST, Dehaene I, Curthoys IS, Todd MJ (1994)
Jerk-waveform see-saw nystagmus due to unilateral mesodiencephalic lesion. Brain 117:789-803
Hedges TR, Hoyt WF (1982) Ocular tilt reaction due to an upper
brainstem lesion: paroxysmal skew deviation, torsion, and
oscillation of the eyes with head tilt. Ann Neuroll1:537-540
Heimchen C, Glasauer S, Bart! K, Bttner U (1996)
Contralesionally beating torsion al nystagmus in a unilateral
rostral midbrain lesion. Neurology 47:482-486
Henn V (1992) Pathophysiology of rapid eye movements in the
horizontal, vertical and torsional directions. In: Bttner U,
Brandt Th (eds) Ocular motor disorders of the brain stern.
Bailliere Tindall, London, pp 373-391
Hess WR, Brgi S, Bcher V (1946) Motorische Funktion des
Tektal- und Tegmentalgebietes. Monatschr Psychiat Neurol
112:1-52
Hrnsten G (1974) Wallenberg's syndrome. Part I. General symptomatology, with special reference to visual disturbances and
imbalance. Acta Neurol Scand 50:434-446
Hyde JE, Toczek S (1962) Functional relation of interstitial nucleus
to rotatory movements evoked from zona incerta stimulation. J
NeurophysioI25:455-466
Keane JR (1975) Ocular skew deviation. Analysis of 100 cases.
Arch NeuroI32:185-190
Keane JR (1985) Alternating skew deviation: 47 patients.
Neurology 35:725-728
King WM, Precht W, Dieringer N (1980) Synaptic organization of
frontal eye field and vestibular afferents to interstitial nucleus
of Cajal in the cat. J NeurophysioI43:912-928
King WM, Fuchs AF, Magnin M (1981) Vertical eye movementrelated responses of neurons in midbrain near interstitial
nucleus of Cajal. J Neurophysiol 46:549-562
Kohler I (1956) Die Methode des Brillenversuches in der
Wahrnehmungspsychologie mit Bemerkungen zur Lehre der
Adaption. Z Exp Angew Psychol 3:381-417
Lang W, Bttner-Ennever J, Bttner U (1979) Vestibular projections to the monkey thalamus: an autoradiographie study. Brain
Res 177:3-17
Larmande P, Missoum A, Tayoro J, Belin C (1994) L'illiusion
visuelle d'obliquite. Rev Neurol (Paris) 150:388-390
Lawden MC, Bronstein AM, Kennard C (1995) Repetitive paroxysmal nystagmus and vertigo. Neurology 45:276-280
Lopez L, Bronstein AM, Gresty MA, Rudge P, Du Boulay EPGH
(1992) Torsional nystagmus: a neuro-otological and MRI study
of35 cases. Brain 115:1107-1124
Lopez L, Ochoa S, Mesropian H, Lacman M, Granillo R (1995)
Acute transient upside-down inversion of vision with brainstem-cerebellar infarction. Neuro-ophthalmology 15:277-280

Vestibular disorders in (frontal) roll plane


Lueck CJ, Hamlyn P, Crawford TJ, Levy IS, Brindley GS, Watkins
ES, Kennard C (1991) A case of ocular tilt reaction and torsional nystagmus due to direct stimulation of the midbrain in man.
Brain 114:2069-2079
Martinez GA, Fox G (1973) Skew deviation. J Ky Med Assoc
17:171-173
Masdeu JC, Gorelick PB (1988) Thalamic astasia: inability to stand
after unilateral thalamic lesions. Ann NeuroI23:596-603
Masdeu JC,Alampur U, Cavaliere R, Tavoulareas G (1994) Astasia
and gait failure with damage of the pontomesencephalic locomotor region. Ann NeuroI35:619-621
Melvill-Jones G (1977) Plasticity in the adult vestibulo-ocular
reflex arc. Phil Trans R Soc Lond 278:319-334
Merfeld DM, Teiwes W, Clarke AH, Scherer H, Young LR (1996)
The dynamic contributions of the otolith organs to human ocular torsion. Exp Brain Res 110:315-321
Mitchell JM, Smith JL, Quencer RM (1981) Periodic alternating
skew deviation. J Clin Neuro-Ophthalmoll:5-8
Morrow MJ, Sharpe A (1988) Torsional nystagmus in the lateral
medullary syndrome. Ann NeuroI24:390-398
Mossman S, Halmagyi GM (1997) Partial ocular tilt reaction due
to unilateral cerebellar lesion. Neurology 49:1-4
Moster ML, Schatz NJ, Savino PJ, Benes S, Bosley TM, Sergott RC
(1988) Alternating skew on lateral gaze (bilateral abducting
hypertropia).Ann NeuroI23:190-192
Noseworthy JH, Ebers GC, Leigh RJ, Dell'Osso LF (1988) Torsional
nystagmus:quantitative features and possible pathogenesis.
Neurology 38:992-994
Ott KH, Kase CS, Ojemann RG, Mohr JP (1974) Cerebellar hemorrhage:diagnosis and treatment. Arch Neurol 31: 160-167
Rabinovitch HE, Sharpe JA, Sylvester TO (1977) The ocular tilt
reaction. A paroxysmal dyskinesia associated with eliptical nystagmus. Arch Ophthalmol 95: l395-1398
Reisine H, Highstein SM (1979) The ascending tract of Deiters'
conveys a head velo city signal to medial rectus motoneurons.
Brain Res 170:172-176
Riordan-Eva P, Faldon M, Bttner-Ennever JA, Gass A, Bronstein
AM, Gresty MA (1996) Abnormalities of torsional fast phase
eye movements in unilateral rostral midbrain disease.
Keurology 47:201-207
Ropper AH (1983) Illusion of tilting of the visual environment.
Report of 5 cases. J Clin Neuro-ophthalmoI3:147-151
Sano K, Sekino H, Tsukamoto Y, Yoshimasu M, Ishijima B (1972)
Stimulation and destruction of the region of the interstitial
nucleus in cases of torticollis and see-saw nystagmus. Confin
NeuroI34:331-338
Schmidt D, KommereIl G (1974) Kongenitaler Schaukelnystagmus
(see-saw nystagmus). Albrecht von Graefes Arch Klin Exp
OphthalmoI191:265-272

197
Schwindt PC, Richter A, Precht W (1973) Short latency utricular
and canal input to ipsilateral abducens motoneurons. Brain Res
60:259-262
Schwindt PC, Precht W, Richter A (1974) Monosynaptic excitatory
and inhibitory pathways from medial midbrain nuclei to
trochlear motoneurons. Exp Brain Res 20:223-238
Silfverskild P (1965) Skew deviation in Wallenberg's syndrome.
Acta Neurol Scan 41:381-386
Slavin ML, Lopinto R (1987) Isolated environmental tilt associated
with lateral medullary compression by dolichoectasia of vertebral artery. J Clin Neuro-ophthalmol 7:29-33
Smith BH (1960) Vestibular disturbance in epilepsy. Neurology
10:465-469
Smith JL, David NJ, Klintworth G (1964) Skew deviation.
Neurology 14:96-105
Solms M, Kaplan-Solms K, Saling M, Miller P (1988) Inverted
vision after frontal lobe disease. Cortex 24:499-509
Steiner I, Shahin R, Melamed E (1987) Acute "upside-down" reversal of vision in transient vertebrobasilar ischaemia. Neurology
37:1965-1966
Suzuki JI, Tokumasu K, Goto K (1969) Eye movements from single
utricular nerve stimulation in the cat. Acta Oto-Laryngol
68:350-362
Szentagothai J (1943) Die zentrale Innervation der
Augenbewegungen.Arch Psychiat Nervenkr 116:721-760
Teuber HL, Mishkin M (1954) Judgement of visual and postural
vertical after brain injury. J PsychoI38:161-175
Tiliket, C, Ventre-Dominey J, Vi ghetto A, Grochowicki M (1996)
Room tilt illusion. A central otolith dysfunction. Arch Neurol
53:1259-1264
Tokumasu K, Suzuki JI, Goto K (1971) A study of the current
spread on electrical stimulation of the individual utricular and
ampullary nerves. Acta Otolaryngol (Stockh) 71 :313-318
Verma AK, Maheshwari MC (1986) Hypaesthetic-ataxia-hemiparesis in thalmic haemorrhage. Stroke 17:49-51
Wakusawa S (1973) Sylvian aqueduct syndrome and mesencephalic skew deviation. Jpn J Ophthalmol17: 154-165
Westheimer G, Blair SM (1975a) The ocular tilt reaction - a brainstern ocular motor routine. Invest OphthalmoI14:833-839
Westheimer G, Blair SM (l975b) Synkinese der Augen- und
Kopfbewegungen bei Hirnstammreizungen am wachen
Macacus-Affen. Exp Brain Res 24:89-95
Wiest G, Baumgartner C, Schnider P, Trattnig S, Deecke L, Mueller
C (1996) Monocular elevation paresis and contralateral
downgaze paresis from unilateral mesodiencephalic infarction.
J Neurol Neursurg Psychiatry60:579-581
Zee DS (1996) Considerations on the mechanisms of alternating
skew deviation in patients with cerebellar lesions. J Vestib Res
6:395-401

Vestibular disorders in (sagittal)


pitch plane

A striking difference between vestibular tone imbalance in the roll and pitch planes is that roll dysfunction is caused by unilateral and pitch dysfunction by
bilateraliesions of paired pathways in the brainstem
or of the cerebellar fiocculus (Brandt and Dieterich
1995). This structural difference probably explains
whya vestibular tone imbalance in pitch frequently
occurs with various intoxications or metabolic disorders. This is unusual for tone imbalance in yaw or
roll, unless as a functional decompensation of an
earlier (compensated) tone imbalance. Downbeat
and upbeat nystagmus are not merely ocular motor
disorders, but central vestibular disorders that also
affect orientation and balance. A tone imbalance in
the pitch plane manifests as vertical upbeating .or
downbeating nystagmus, fore-aft head-and-body tIlt,
and deviation of the subjective horizontal toward the
direction of the slow phase of the nystagmus.
Clinically, downbeat nystagmus occurs more frequently than upbeat nystagmus and is often permanent (as in Arnold-Chiari malformation),
whereas upbeat nystagmus is usually a transient
phenomenon. Lesional sites for upbeat nystaw~1Us
have been more precisely confirmed by chmcal
studies (bilateraliesions of the pontomesencephalic
junction or the medulla) than those for d~wnbeat
nystagmus (bilateral pontomedullary leslOns or
bilateral flocculus dysfunction). In contrast, the
pathomechanism of downbeat nystagmus appears to
be dearer (tone imbalance of the VOR in pitch) than
that of upbeat nystagmus, which can result from
different pathomechanisms: a (vestibular?) pontomesencephalic and a (non-vestibular?) medullary
one. Transitions behveen upbeat and downbeat
nystagmus have been frequently described in paramedian pontomedullary lesions. Both types occur
with paramedian plaques in multiple sderosis,
cerebellar degeneration, or drug intoxication. While
downbeat nystagmus is more typical for congenital
cervical malformations (Arnold-Chiari malformation), upbeat nystagmus is more typical for bilateral
brainstern ischaemia (basilar artery thrombosis) or
brainstem tumours. Upbeat and downbeat nystag-

mus in the primary position of gaze may be the


result of various intoxications (without structural
lesion). Management can either successfully dampen
the nystagmus (and alleviate oscillopsia) or worsen
it, depending on the drugs used.

Downbeat nystagmus
(vestibular downbeat syndrome)
Downbeat nystagmus in the "primary" gaze position,
or more particularly on lateral gaze, is often accompanied by oscillopsia and postural instability. This is a
dearly defined and, depending on the lesional site,
permanent association of symptoms, which often
indicates structural lesions of the para median
craniocervical junction (Cogan 1968). It involves eyehead coordination in the pitch plane, mediated by
"ascending" pathways from the vertical semicircular
canals and the otoliths in functional cooperation with
"descending" pathways from the rostral midbrain
integration centre for the pitch plane. Downbeat nystagmus is not a purely ocular motor disorder, but a
central vestibular disorder that affects perception and
balance. Differential diagnoses indude gaze-evoked
nystagmus, acquired pendular nystagmus, spasmus
nutans and rare vertical forms of congenital nystagmus; it must not be confused with ocular bobbing in
the comatose patient. Downbeat nystagmus differs
from these other conditions in that it is a spontaneous
jerk nystagmus, wh ich is activated by lateral gaze and
is not suppressed by fixation. It can be successfully
treated with GABA-agonists, such as baclofen, or by
surgical decompression in cases of Arnold-Chiari
malformation.

The clinical syndrome


The clinical ocula-r motor abnormalities have been
frequently described (Cogan 1968; Baloh and
Spooner 1981; Halmagyi et al 1983).
199

200

Vertigo

Nystagmus

Downbeat nystagmus is present in darkness as weIl


as with fixation; slow-phase velocity and amplitude
increase on lateral gaze (Fig. 11.1) or with head
extension (Figs 11.2, 11.3) or head movements in the
sagittal (pitch) plane. Downbeat nystagmus may be
present only on downward or lateral gaze. Slow
phase velo city is not consistently related to vertical
gaze and, contrary to Alexander's law, may even be
maximal on upward rather than downward gaze.
Nystagmus is a jerk, usually with linear slow phases.
It may exhibit changes of exponential velo city in
slow phases, both increasing and decreasing (Abel et
al. 1983; Lavin et al. 1983). It has been speculated that
these are short-term gain changes produced by cerebellar compensation for leaky brainstem neural
integrators. Convergence can enhance or dampen
the nystagmus in some patients; loss of fixation
(lights extinguished) either does not affect the nystagmus or slightly decreases frequency with a simultaneous increase in amplitude. Downbeat nystagmus
may be episodic in Arnold-Chiari malformation (Yee
et al. 1983) or paroxysmal (Lawden et al. 1995).
It has been reported that reversals from downbeat
to upbeat nystagmus can be provoked by upward
gaze deviation (Baloh and Spooner 1981), convergence (Cox et al. 1981), or transitions from sitting to
supine position (Crevits and Reynaert 1991).
Downbeat and upbeat nystagmus are the directional
counterparts of a vestibular tone imbalance in the
pitch plane. The dose proximity of the areas causing
either upbeat or downbeat nystagmus in the medulla
agrees with the directional changes between the two.
Reversals from upbeat to downbeat nystagmus have
also been observed (Cox et al. 1981; Rousseaux et al.

"
Z

IU

,.

40 left

'ii

0
CI.

Oscillopsia and impaired motion perception

The patients complain of a distressing illusory oscillation of the visual scene (oscillopsia, p. 430) and postural imbalance. Both are obligatory but hitherto
poorly studied symptoms of the syndrome. The
retinal slip in downbeat nystagmus is misinterpreted
as motion of the visual scene, because the involuntary
ocular movements that override fixation are not associated with an appropriate efference-copy signal.
Oscillopsia is a permanent symptom, but the illusory
motion is less than would be expected from the
amplitude of the nystagmus; it increases with increas-

.tr.lght .h d

2~.~ ~

40 rlght
~

5.

20
d

... 10N:~

1991; Sakuma et al. 1996). The description of a


monocular downbeat nystagmus (Bogousslavsky
and Regli 1985) in cerebellar degeneration is difficult
to interpret, since dissociation of eye movements
suggest inter- or infranudear dysfunction. We have
observed transient monocular downbeat nystagmus
of the left eye in a patient with a right thalamic
haemorrhage and contraversive ocular tilt reaction
(Brandt 1991). Mochizuki et al. (1996) described a
similar patient who presented with ipsilateral oculomotor nerve palsy and contralateral downbeat
nystagmus due to paramedian thalamopeduncular
infarction. They attribute the monocular downbeat
nystagmus to the palsy of the other eye.
Horizontal and vertical smooth pursuit and optokinetic nystagmus are impaired, particularly downward pursuit. The downward pursuit defect per se
does not cause the nystagmus, but pursuit asymmetry may be the consequence (superimposition) of
spontaneous nystagmus (Daroff et al. 1973). The
same is true for the impaired ability of these patients
to suppress vestibular nystagmus by fixation.

~
I

~
5.

>---------<

Fig.ll.l. Simultaneous recording of vertical eye movements and body sway in a patient with down beat nystagmus/vertigo syndrome during unsupported stance with the head upright and fixating a target either straight ahead or 40 laterally. Downbeat nystagmus is activated during lateral gaze, which also increases body sway amplitudes, especially in the fore-aft direction.

Vestibular disorders in (sagittal) pitch plane

201

ENG

verlieal

posll6al
sway
lateral

Eyes closed
poall6a1
sway

1OHm
1OHm

Fig. 11.2. Simultaneous recording of vertical eye movements and body sway during unsupported stance with the head either
upright or extended. Postural instability in down beat nystagmus is not simply due to ocular vertigo because there is marked fore-aft
ataxia even with the eyes closed. This becomes particularly apparent during head extension (bottom). Moreover, with the eyes open,
visual control of balance is discernible but somewhat diminished by concurrent activation of the downbeat nystagmus by head extension (top).

ing amplitude. The mean ratio between the two is


0.37 (Bchele et al. 1983). Motion perception is
impaired in patients with acquired downbeat
nystagmus (Bchele et al. 1983; Dieterich and Brandt
1987), particularly in the vertical plane in the direction of slow phases (Dieterich et al. 1998). A reduced
sensitivity of motion perception is beneficial to these
patients if it also alleviates the disturbing oscillopsia.
Visual orientation is impaired in the vertical
plane: tilts of the visual coordinates (5-8 deg) are
perceived as being opposite in direction to the fast
phases of the acquired nystagmus (Dieterich et al.
1998). Both the disturbance of spatiallocalisation
and the oscillopsia are not restricted to the fovea but
involve the entire visual field, and therefore affect
the two modes of visual processing, "focal" and
"ambient", respectively.

Postural imbalance
Oscillopsia should be expected to cause an impairment of postural balance, since retinal image motion
is a major cue for body stabilisation. However, this
kind of "visual ataxia" cannot simply ac count for the
typical postural imbalance, which is a strikingfeature of the fore-aft body sway and includes a tendency to fall backwards (Figs 11.2, 11.3). This
fore-aft postural instability can be interpreted to be

a direction-specific vestibulospinal (or cerebellar)


imbalance, since it can be observed when the eyes
are closed. We believe that the objective measurable
backward tilt represents a vestibulospinal compensation in the direction opposite to the perceived
lesional "forward vertigo;' which corresponds to
downbeat nystagmus (Bchele et al. 1983; Brandt et
al. 1986). When the eyes are open, a measurable
visual stabilisation of body sway is preserved, but it
does not sufficiently compensate for the visual ataxia
(Fig. 11.3). In downbeat nystagmus (more aptly
termed "vestibular downbeat syndrome"), the
patient's pathological postural sway with the eyes
open depends on the direction of gaze; it increases
with increasing amplitude of the nystagmus (Figs
11.1, 11.3). Pathophysiologically it is secondary to a
combination of both vestibulospinal ataxia and
reduced visual stabilisation of posture owing to the
nystagmus.

Aetiology and pathomechanism


Pathomechanism and site of the lesions
The original and appealing proposal that there is a
downward pursuit defect assumed that downward
velocity information cannot be transmitted to the

202

Vertigo

fore- aft body sway

SR
4

pat ients

z
'"'"

.."
>
~

'"g-"
c

'"'"

n ~ 6

E 0

eyes closed

2
3

latera l body sway

Fig.11.3. Posturography of the fore-aft and lateral body sway


(root mean square values, Newton metres) in six patients with
down beat nystagmus {shaded columnsJ compared to a control
group of normal individuals. Vestibulocerebellar ataxia is manifest in the abnormal fore-aft body sway with the eyes cIosed .
Impaired visual control of postural balance is particularly apparent on lateral gaze, which increases the nystagmus intensity.
(From Bchele et al. 1983.)

neural integrator that generates smooth eye movements in response to commands from the visual
system (Zee et al. 1974). This is no longer acceptable,
since it cannot explain all the features of the syndrome, far example, ataxia. It was questioned by
Baloh and Spooner (1981), who assurne "that downbeat nystagmus is a type of central vestibular nystagmus resulting from an imbalance in the central
vertical vestibulo-ocular pathways, due either to a
lesion in the floar of the fourth ventricle between the
vestibular nuclei (which interrupts the tonic excitatory activity to the inferior recti), or to abilateral
lesion of the flocculus (which leads to an increase in
tonic excitatory activity to the superior recti due to a
disinhibition)". Both lesions (Fig. 11.4) have been
shown to cause downbeat nystagmus in animals
(DeJong et al. 1980; Takemori and Suzuki 1977; Zee
et al. 1981), but clinical case studies are stillless preeise as regards the causative structures. Rare case
descriptions of a monocular downbeat nystagmus
with acute mesodiencephalic or thalamopeduncular
lesion offer a third critical site (Jacome 1986; Brandt
1991; Mochizuki et al. 1996).
Since head tilt in a stationary patient modulates

Nu -'

",. . . . . v

Fig.11.4. Mechanism of downbeat nystagmus resulting from


an imbalance in the central vertical vestibulo-ocular pathways
due to lesions either between the vestibular nucJei (which interrupt tonic excitatory activity to the inferior recti) or the flocculus
(disinhibition in tonic excitatory activity to the superior recti) as
hypothesised by Baloh and Spooner (1981). SR, superior rectus;
IR, inferior rectus; AC, anterior canal; HC, horizontal canal; PC, posterior canal; VN, vestibular nucJei; Be, brachium conjunctivum;
MLF, medial longitudinal fasciculus; 111, oculomotor complex; IV,
trochlear nucleus; VI, abducens nucJeus; U, uvula; T, tonsil; PF,
paraflocculus; N, nodulus; F, flocculus;V, visual input; Nu, nuchal
input. Dashed fines indicate multisynaptic pathways. Darkened
neuron is inhibitory.

the intensity of the nystagmus, otolith function is


probably involved in addition to the semicircular
canals. Downbeat nystagmus may arise from asymmetrie vertical canal reflexes whose gains are modulated by otolithic influences (Gresty et al. 1986).
It is not yet clear what causes transition from
downbeat to upbeat or, vi ce versa, upbeat to downbeat nystagmus, especially in paramedian pontomedullary lesions. Wh at determines the direction of
vertical imbalance in pitch plane? The site of the
particular lesion involving either pathways far up or
down VOR? The differential effects of vestibulocerebellar inhibition on vertical VOR? The difference
in excitatory and inhibitory neural transmission of
the VOR in pitch? That downbeat is more frequent
than upbeat nystagmus may be related to a physiological pitch plane asymmetry with a tendency of
upward ocular drift in darkness and increase in head

Vestibular disorders in (sagittal) piteh plane

tilt from upright as demonstrated in the eat (Rude


and Baker 1996). The latter eorresponds to earlier
observations of a eomparable asymmetry of optokinetie nystagmus in animals (Matsuo and Cohen
1984) and humans (Takahashi et al. 1987). This
asymmetry disappears with ablation of the maeular
end-organs in the monkey (Igarashi et al. 1978) and
in humans who adapt to mierogravity during spaee
missions (Clement et al. 1986).
Downbeat nystagmus, when eaused by alesion
between the vestibular nuclei, is usually a permanent
syndrome, sinee it involves the eommissural fibres
neeessary for eentral eompensation of a vestibular
tone imbalance (Brandt et al. 1986).
Aeriology

The two most common causes of a downbeat nystagmus/vertigo syndrome are cerebellar ectopia (25%)
(Arnold-Chiari malformation, for example, Fig. 11.5)
and cerebellar degeneration (25%), e.g. olivopontocerebellar degeneration. A further 10-20% of patients
have a variety of conditions (Cogan 1968; Halmagyi et
al. 1983), and in about 30% an unequivocal diagnosis
of the cause cannot be established.
In cerebellar degeneration and drug-induced
downbeat nystagmus, an asymmetrie vestibulocerebellar disinhibition of the "Purkinje cell activity" on
the vertical canal reflexes may be causative.
Nutritional cerebellar syndromes due to thiamine
deficiency, in particular alcoholic cerebellar degeneration (Fig. 11.6), not only cause a typical 3 Hz fore-

Fig.11.5. Cerebellar ectopia in a patient suffering from downbeat nystagmus due to Arnold-Chiari malformation I.

203

aft oscillation of body sway (Dichgans et al. 1976)


but also downbeating nystagmus (Cogan 1968;
Costin et al. 1980; Zasorin and Baloh 1984).
Antiepileptic drugs, especially phenytoin (Alpert
1978) and carbamazepine (Sullivan et al. 1981;
Wheeler et al. 1982), can produce a reversible downbeat nystagmus with associated cerebellar signs,
depending on the dosage of the drugs. Other causes
include lithium toxicity (Halmagyi et al. 1983, 1989;
Coppeto et al. 1983; Corbett et al. 1989), felbamate
intoxication (Hwang et al. 1995), or toluene abuse
(Malm and Lyiug-Tune1l1980). Severe magnesium
depletion (Saul and Selhorst 1981) and vitamin B12
deficiency (Mayfrank and Thoden, 1986) have been
reported to result in downbeat nystagmus.
Other conditions associated with downbeat nystagmus are multiple sclerosis (Bronstein et al. 1987;
p. 244), familial episodic ataxia (p. 365), tumours of
the posterior fossa, cerebellar degeneration (Baloh
and Yee 1989), paraneoplastic cerebellar degeneration (Hammack et al. 1992), infratentorial vascular
diseases such as dolichoectasia of the vertebrobasilar artery (Jacobson and Corbett 1989),
haematomas, cavernomas, syringobulbia (Bertholon
et al. 1993) and encephalitis (see Table 11.1).
Sometimes downbeat nystagmus is lithium
induced, e.g. in pre-existing Arnold-Chiari malformation (Monteiro and Sampaio 1993), or caused as

Fig. 11.6. Alcoholic cerebellar degeneration (most pronounced in the anterior lobe of the cerebellum) in a patient suffering from down beat nystagmus.

204

Vertigo

Table 11.1. Aetiology of vertical nystagmus

Table 11.2.

A Downbeat nystagmus

Clinicalsyndrome
- Downbeat nystagmus in the primary position of gaze (no suppression
by fixation). increased on lateral gaze or head extension
- Associated distressing oscillopsia and postural imbalance with a
tendency to fall backward
- Saccadic downward pursuit
- Transitions from down beat to upbeat nystagmus possible

Cerebellar degeneration, including familial periodic ataxia, and


paraneoplastic degeneration
Craniocervical anomalies, including Arnold-Chiari malformation, Paget's
disease, basilar invagination
Infarction of brainstem or cerebellum
Dolichoectasia ofthe vertebrobasilar artery
Multiple sclerosis
Cerebellar tumour, including haemangioblastoma
Syringobulbia
Encephalitis
Head trauma
Anticonvulsant medication
Lithium intoxication
Alcohol, including cerebellar degeneration
Wernicke's encephalopathy
Magnesium depletion
Vitamin B12 deficiency
Toluene abuse
Congenital
Midbrain infarction
Increased intracranial pressure and hydrocephalus
Transient finding in otherwise normal infants
B Upbeat nystagmus
Cerebellar degenerations and atrophies
Multiple sclerosis
Infarction of medulla or cerebellum
Tumours ofthe medulla, cerebellum, or midbrain
Wernicke's encephalopathy
Brainstem encephalitis

Downbeat nystagmus (vestibular down beat syndrome)

Incidence/age/sex
Depends on aetiology, no obvious preference of sex, rare in children
(congenital)
Pathomechanism
Tone imbalance of the vertical semicircular canal reflexes (pitch plane)
modulated by otolithic input
Structural or functionallesions involve
- either the floor of the fourth ventricle between the vestibular nuclei
- orvestibulocerebellar flocculus (intoxication, cerebellar
degeneration)
Aetiology
The two most common causes are cerebellar ectopia and cerebellar
degeneration including alcoholic cerebellar degeneration
Other conditions: drugs (phenytoin, carbamazepine, lithium). multiple
sclerosis, tumour, haematoma, vascular disease, encephalitis,
magnesium depletion, vitamin B12 deficiency (see Table 11.1)
Course/prognosis
Frequently permanent when caused by structurallesions
Usually reversible when caused by intoxication or metabolic deficiency
Management
Depends on aetiology, Le. surgical decompression
Medical treatment with baclofen (or clonazepam?)
Differential diagnosis
Acquired pendular nystagmus, gaze-evoked nystagmus, upbeat
nystagmus, spasmus nutans (infants), vertical congenital nystagmus,
ocular bobbing

Beh~et's syndrome
Meningitis

Leber's congenital amaurosis or other congenital disorder of the anterior


visual pathways

Management

Thalamic arteriovenous malformation

Downbeat nystagmus due to drugs, magnesium


depletion, or vitamin B12 deficiency is usually
reversible when the intoxication or the metabolie
deficiency is reversed. Downbeat nystagmus due to
structural lesions in the posterior fossa is usually
permanent, although a surgical suboccipital decompression in Arnold-Chiari malformation to relieve
the compression of the herniating cerebellum
against the caudal brainstem may lead to gradual
improvement of some of the distressing symptoms
(Spooner and Baloh 1981). Suboccipital craniotomy
(Pedersen et al. 1980), transoral removal of the
odontoid process in the basilar impression (Senelick
1981) or of an osteophyte compressing the vertebral
artery (Rosengart et al. 1993), and surgical decompression of a syringomyelic cyst in the medulla
(Pinel et al. 1987; Chan et al. 1991) were able to
resolve the syndrome in individual patients. For one

Congenital
Organophosphate poisoning
Tobacco (nicotine)
Associated with middle ear disease
Transient finding in otherwise normal infants
(From Leigh and Zee 1991)

an intermittent syndrome by head tilt due to vertebral artery compression (Rosengart et al. 1993) or by
a vermian arachnoid cyst with associated obstructive hydrocephalus (Chan et al. 1991).
Downbeat nystagmus is rare in children. It may be
congenital hereditary (Bixenman 1983) as a persisting syndrome, or it may occur during infancy and
resolve naturally (Weissman et al. 1988).

205

Vestibular disorders in (sagittal) pitch plane

patient, base-out prisms were added to both spectade lenses, because the convergence both dampened the nystagmus and decreased the oscillopsia
(Lavin et al. 1983).
Target symptoms for symptomatic medical treatment are distressing oscillopsia and reduced visual
acuity owing to the fixation nystagmus. Postural
imbalance is less prominent and less distressing. It is
our experience that badofen suppresses downbeat as
well as upbeat nystagmus and associated postural
imbalance in most cases (Fig. 11.7; Dieterich et al.
1991). Badofen appears to have a GABA-B-ergic
effect and to augment the physiological inhibitory
influence of the vestibulocerebellum on the vestibular nudei. It was previously recommended for treatment of periodic alternating nystagmus (Halmagyi
et al. 1980). Muscarinic antagonists, in particular the
anticholinergic drug scopolamine, reduced nystagmus in five patients with acquired pendular nystagmus and in two patients with downbeat nystagmus
(Barton et al. 1994). Benztropine was less effective.
The GABA-A agonist donazepam (3 x 0.5 mg po
daily) was used with some success to treat nystagmus and oscillopsia (Chambers et al. 1983; Currie
and Matsuo 1986; McConnell et al. 1990). Physical
balance training improves postural instability in
patients with a newly acquired downbeat nystagmus
syndrome (Brandt et al. 1986). Table 11.2 summarises the information given in this chapter about the
downbeat nystagmus/vertigo syndrome.

Upbeat nystagmus (vestibular upbeat


syndrome)
Upbeat nystagmus in the primary position of gaze
with concomitant oscillopsia and postural instability
is a pendant of downbeat nystagmus, and most probably reflects an imbalance of vertical vestibuloocular reflex tone. It has the same causes and
involves central eye-head co ordination in the pitch
plane as mediated by pathways from the vertical
semicircular canals and the otoliths. Since the manifestations are typically modulated by otolithic input
arising from static head tilt (Fig. 11.8), upbeat
nystagmus is in a broader sense also a kind of positional nystagmus (see central positional nystagmus,
p. 292). As distinct from downbeat nystagmus, brainstern lesions are often found in patients with upbeat
nystagmus. Two separate intra-axial brainstem
lesions in the tegmentum of the pontomesencephalic
junction and in the medulla near the perihypoglossal nudei (Fig. 11.9) are likely to be responsible for
this syndrome, but there is insufficient evidence to
determine whether the cerebellar vermis is involved.
Upbeat nystagmus indicates bilaterallesions of the
pathways mediating VOR in pitch. However, not all
forms of upbeat nystagmus must be vestibular. It can
also arise from acquired asymmetries of the smooth
pursuit or the gaze-holding neural integrator system.

betor. bacloten

;]
;]
U

with badoten

;]
;3
U

eyes closed

fixalion. straight ahead

:B'

40" 18ft

40" right

2~]
20"

Fig.ll.7.

;3
0

20"""
20" down

;J
U

Partial suppression of downbeat nystagmus with badofen in a patient suffering from multiple sderosis.Original recordings of vertical ENG.

Vertigo

206

Differential diagnosis includes gaze-evoked nystagmus, convergence-evoked nystagmus, acquired pendular nystagmus, spasmus nut ans and rare forms of
vertical congenital nystagmus. It must not be confused with reversed ocular bobbing in comatose
patients. Since upbeat nystagmus can be caused by
metabolie or pharmacological intoxication (nicotine, antiepileptic drugs), it can be effectively suppressed by, e.g. GABAergic drugs.

head upright, sitting

ENG

20'

vertieal

head extension. supine

The clinical syndrome

5.

Fig.11.8. Original electronystagmographic recording of vertical eye movements in a patient with upbeat nystagmus. The
nystagmus is dampened by static head tilt to the supine position
(bottom).

G 0
'(

midbraln

~
BC

pons

I
I
I

VT

medulla

t l t 7'
PHN

BCVT

Schematic representation of the three-neuron excitatory reflex arc from the anterior semicircular canal (AC), the superior
vestibular nucleus (VIII), the brachium conjunctivum (BC) to the
contra lateral oculomotor nucleus (111), the superior rectus muscle
(SR) and inferior oblique muscle (10). Also shown is the excitatory
ventral tegmental pathway (VT), which connects the superior
vestibular nucleus with the contralateral oculomotor nucleus. An
ascending pathway from the perihypoglossal nuclei (PHN) in the
medulla possibly modulates the tone of the vertical VOR. The
arrows indicate the lesions of the brachium conjunctivum, the ventral tegmental pathway,and the perihypoglossal connection wh ich
may produce upbeat nystagmus if affected bilaterally (only unilateral pathways are depicted for the sake of clarity).
Fig.11.9.

Nystagmus
Upbeat nystagmus with fixation was first described
by Stengel (1935). It is a jerk nystagmus and, in most
cases, obeys Alexander's law, being greatest in
upgaze. Slow phases have linear or decaying wave
forms. Unlike downbeat nystagmus, it may not be
affected by lateral gaze or the effects may be variable
(Fisher et al. 1983).
Convergence can exaggerate or dampen upbeat
nystagmus (Fisher et al. 1983) or even convert the
nystagmus to downbeating (Cox et al. 1981; Carl et al.
1982; Rousseaux et al. 1991), a complex and hitherto
unexplained interaction between the voluntary and
reflex oculomotor systems. In acquired convergenceevoked nystagmus the vertical component is most
common, with upbeat being more frequent than
downbeat (Oliva and Rosenberg 1990). In darkness,
upbeat nystagmus can be partially suppressed, or the
eyes can show a downward drift.
Static head tilt to the prone and supine positions
modifies the characteristics of the nystagmus in
most cases (Fisher et al. 1983): nystagmus may be
enhanced, suppressed (Fig. 11.8), or even reversed in
direction to downbeat (Mizuno et al. 1990).
All patients exhibit saccadic upward pursuit
(owing to the nystagmus?) with corresponding
deficits in optokinetic nystagmus, but some of them
have additional downward pursuit defects. The vertical
vestibulo-ocular reflex, however, appears basically
intact, since slow phase compensatory eye movements can be elicited by up-and-down head movements in the pitch plane. The tone imbalance could
account for the obvious asymmetry.

Oscillopsia, motion perception and spatial


orientation
Because of the involuntary retinal slip of the visual
scene, patients complain of a distressing oscillopsia
and reduced visual acuity, as do patients with downbeating nystagmus (p. 200). Oscillopsia is always
smaller than the net retinal slip due to impaired

207

Vestibular disorders in (sagittal) piteh plane

motion perception with acquired ocular oscillation


(Bchele et al. 1983; Dieterich and Brandt 1987).
Visual motion pereeption is predominantly
impaired in the vertical plane, and psychophysieal
adjustments of the subjective horizontal exhibit significant downward deviations (Dieterich et al. 1998).

Postural imbalance
Postural imbalance in the fore-aft direction is another
manifestation of the syndrome, whieh has not yet
been thoroughly studied. By analogy with other
vestibular vertigo syndromes, one would expeet a
tendency to fall forward (motor compensation of an
apparent backward tilt). Some of our patients, however, behaved differently: they showed a tendency to
fall backward, i.e. in the same direction as patients
with downbeat nystagmus. A decreased sensitivity
for the perception of body verticality (subjeetive
postural vertieal) has been reported for patients
with upbeat or downbeat nystagmus (Bisdorff et al.
1996). Some patients (non-vestibular upbeat nystagmus?) do not seem to have postural instability.

Aetiology and pathomechanism


Pathomechanism and site of the lesions
The pathways for downward and upward vestibuloocular reflexes run separately from the vestibular
nuclei to the oculomotor nuclei. Upbeat nystagmus
is assumed to arise from a lesional or functional tone
imbalance, which is due to disruption of the upward
VOR pathways within the brachium conjunctivum
(Benjamin et al. 1986), the ventral tegmental pathway (Ranalli and Sharpe 1988), or of struetures of
the caudal brainstem which modulate VOR in pitch,
e.g. the perihypoglossal nuclei (Nakada and Remler
1981; Keane and Itabashi 1987). We agree with Leigh
and Zee (1991) that acquired vertical fixation nystagmus may involve not only the VOR but also the
smooth pursuit and gaze-holding system: "Common
to all hypotheses for downbeat and upbeat nystagmus is the faet that, unlike the horizontal vestibular
system, whieh is right-Ieft symmetric, the anatomie
connections for vertical vestibular responses are dissimilar for upward or downward eye movements.
These up-down asymmetries may involve connections necessary for VOR, the otolith-ocular reflexes,
the vestibulocerebellum, gaze-holding network
(neural integrator) and the smooth-pursuit system."
Vestibular and non-vestibular upbeat nystagmus
syndromes may exist, which have not yet been clearly
differentiated either clinically or with respect to the
lesional sites.

The pathways and neural transmitters mediating


the up-and-down tone in the piteh plane are different. This explains why the causative lesions are
separate and why certain toxic effects such as
nicotine-indueed upbeat nystagmus are directionally
specific. On the other hand, transitions from upbeat
to downbeat nystagmus or downbeat to upbeat nystagmus with head tilt, eonvergence, or simply during
the course of the disease reflect the intimate network
that maintains balance in the pitch plane. As a rule,
downbeat and upbeat nystagmus are elicited only by
bilaterallesions. Theoretieally, the VOR in pitch can
be mediated by the same "gravieeptive" pathways as
the VOR in roll (Brandt and Dieterich 1995; p. 167).
Then the vestibular system may change its functional
plane of action from roll to pitch by switehing from
a unilateral to a bilateral mode of operation (Fig.
C.4). Accordingly, upbeat nystagmus can be explained by bilaterallesions of the crossed VOR pathways of the anterior semicireular canal (Fig. 11.10). If
bilaterallesions are asymmetrie, a combination of
vestibular syndromes in pitch and roll planes would
be expected. In fact, single cases with lateropulsion
and upbeat nystagmus (Benjamin et al. 1986) or a
spontaneous 300 -right tilt, which suppressed upbeating nystagmus in a haemorrhagic lesion of the left

RIGHT LABYRINTH

LEFT LABYRINTH

paNS

MESENCEPHALON

<A\ Lii>

PITCH

Fig. 11.10. A schematic representation of crossed-excitatory


(anterior semicircular canal: AC) pathways, which convey a mixed
vertical and torsional ocular motor impulse to both eyes (both
eyes upward). A bilaterallesion causes a transition of the plane of
action from roll to pitch, because torsional components cancel
each other, and the 1055 of upward impulse leads to a downward
deviation ofthe eyes. Downward deviation with upward refixation saccades manifests as upbeat nystagmus. (Compare Fig. C.4,
Downbeat nystagmus).

Vertigo

208

brachium conjunctivum and the anterior vermis


(Kattah and Dagi 1990), have been described.
Upward VOR is mediated by excitatory input from
the anterior semicircular canal to the superior
vestibular nuclei and to the contralateral oculomotor
nucleus in the midbrain via the brachium conjunctivum (Ito et al. 1976). Bilateral disconnection of this
ascending route (brachium conjunctivum) would
result in a tone imbalance of vertical gaze, i.e. downward drift of the eyes with compensatory upward
saccades: upbeating nystagmus (Fig. 11.9). As the site
of the lesion in the majority of pathologically verified brainstem lesions associated with upbeat
nystagmus was remote from the brachium conjunctivum, Ranalli and Sharpe (1988) speculated about
the functional significance of a ventral tegmental
pathway, which was identified in cats by Carpenter
and Cowie (1985). This pathway connects the superior
vestibular nucleus with the superior rectus and
inferior oblique subdivision of the contralateral
oculomotor nucleus. Bilateral damage of this ventral
tegmental pathway (if it exists in humans) along the
medial border of the laterallemniscus could account
for the upbeat nystagmus reported by others (Keane
and Itabashi 1987; Fisher et al. 1983; Gilman et al.
1977; Hirose et al. 1991). Fisher and co-workers
(1983) were the first to relate upbeat nystagmus to
two distinct intra-axial brainstem lesions of the
tegmentum, either at the pontomesencephalic or at
the pontomedullary border. The latter has now been
more precisely located in the medulla.
The pontomesencephalic cases may be caused by
lesions of the upward VOR route within the brachium
conjunctivum or the ventral tegmental pathway (Fig.
11.9). They are sometimes associated with internuclear ophthalmoplegia (Troost et al. 1980)
(Fig.11.11). The medullary cases most probably
involve the prepositus hypoglossi nucleus (Fig.
11.12), which receives vestibular input and excites
various oculomotor neurons via the mediallongitudinal fasciculus (Baker and Berthoz 1975),
predominantly the superior rectus and inferior
oblique in the case of vertical eye movements (Baker
and Berthoz 1976). This is supported by animal
experiments in which bilateral electrocoagulation of
the prepositus nucleus resulted in primary position
upbeating nystagmus in the cat (Kato et al. 1985;
Harada et al. 1994). The clinical cases of O'Brian and
Bender (1945) and Kato et al. (1985) showed tongue
deviation, fasciculation, atrophy and hiccups, which
makes alesion around the hypoglossal nucleus likely. Munro et al. (1993) speculated about the possible
role of the nucleus intercalatus, the most caudal of
the perihypoglossal nuclei (a kind of vertical
integrator?), whereas Bttner et al. (1995) discussed
a cell group of paramedian tract neurons ventral to

Fig.11.11. eT scan of a patient with a tumour of the pontomesencephalic tegmentum who suffered from a combination of
upbeat nystagmus with bilateral internuclear ophthalmoplegia.

the caudal nucleus praepositus hypoglossi, which


animal experiments suggest may be involved in vertical gaze holding. The modulation of upbeat nystagmus by static head tilt can only be explained by
otolithic infiuences. It is well known that there is a
convergence of otolith and canal input, and that the
otoliths contribute to vertical and rotatory eye
movements (Precht et al. 1979).
Earlier suggestions that there are two types of
upbeat nystagmus, a large-amplitude type with
anterior cerebellar vermis lesion and a smallamplitude type with medullary lesion (Daroff and
Troost 1973), were not confirmed. There are some
reports of primary position upbeat nystagmus, secondary to experimental (Spiegel and Scala 1941) or
clinical (Bender and Gormen 1949; Pawl 1973)
lesions of the cerebellar vermis (Furman et al. 1986),
but this was not seen in Dandy-Walker syndrome
with vermis aplasia or after experimental ablation of
the vermis in the monkey (Blair and Gavin 1979).
Further, the involvement of the inferior olive, a
lesion of which was reported to induce vertical
nystagmus (Murphy and O'Leary 1971), was not confirmed by others (Haddad et a1.l980).
Aetiology

Upbeat nystagmus was observed in 26 of 17 900


patients examined at a neuro-otological clinic in

209

Vestibular disorders in (sagittal) pitch plane

Fig.11.12. MR scan of a patient with a haemorrhage due to medullary cavernoma. The patient had upbeat nystagmus which gradually improved after a few weeks.

Japan. The incidence rate was 0.145% (Tokumaso et


al. 1991). The aetiology of upbeat nystagmus is in
general similar to that of downbeat nystagmus
(Table 11.1). Malformation of the craniocervical
junction and cerebellar degeneration seem to be less
frequent than in downbeat nystagmus, whereas
brainstem tumours are more frequent. Upbeat
nystagmus can be associated with bilateral vascular
brainstem lesions (basilar artery thrombosis),
haematoma, cavernoma, multiple sclerosis, encephalitis, abscess, or head injury. It has been repeatedly
reported in alcoholic degeneration, especially in
Wernicke's encephalopathy (Bender and Gormen
1949; Schmidt 1972; Cox et al. 1981; Zumstein and
Meienberg 1982) or in single cases of Fisher's syndrome (Yamazaki et al. 1994), central diabetes
insipidus (Fujikane et al. 1992), Pelizaeus-Merzbacher
disease (Trobe et al. 1991), and even associated with
middle ear disease (Gresty et al. 1988). We have seen
transient upbeat nystagmus associated with various
intoxications, for example with antiepileptic drugs.
Upbeat nystagmus can on rare occasions be congenital (Forsythe 1955; Sogg and Hoyt 1962; Shibasaki et
al. 1978; Hoyt and Gelbert 1984).
Neveling and Kruse (1961) were the first to
describe a tobacco-induced upbeat nystagmus,
which Sibony et al. (1987) investigated more
thoroughly. This nicotine-induced nystagmus (after
smoking one cigarette) can also be elicited by chewing nicotine gum (Sibony et al. 1990). It obeys
Alexander's law, but only occurs in darkness and is
suppressed by visual fixation. It was reported to have

a latency of 40 to 90 seconds, a duration of 10 to 20


minutes and maximum slow-phase velocities at 2 to
3 minutes (Sibony et al. 1987). The degree of impairment in upward pursuit correlated with the intensity
of nicotine-induced nystagmus recorded in darkness
(Sibony et al. 1988). Nicotinic receptors have been
described in different parts of the vestibular endorgan (Hiel et al. 1996), the lateral and medial
vestibular nucleus (Phelan and Gallagher 1992) and
in efferent vestibular neurons (Ishiyama et al. 1995).
It is not yet clear which structures are involved in
nicotine-induced upbeat nystagmus.

Management
Upbeat nystagmus may be associated with severe
vertigo, ataxia and nausea, particularly at first. Such
patients may require vestibular sedatives, e.g. dimenhydrinate or scopolamine, as long as nausea lasts.
Depending on the aetiology, the natural his tory of
this sign usually shows gradual improvement or it
disappears, in contrast to downbeat nystagmus,
which is frequently permanent. Physical exercise
involving fixation, eye movements and postural balance will accelerate central compensation. Medical
treatment is possible with baclofen (3 x 5-10 mg po
daily), which has a beneficial effect on nystagmus
amplitude, oscillopsia and visual acuity in some
patients (Fig. 11.13; Dieterich et al. 1991).
Carbamazepine was found to be effective in a single
case of upbeat nystagmus due to multiple sclerosis
(Iwata et al. 1996).

210

Vertigo

beto.. botlo!..,

20.

Gy" cIoaed

O' ]~

20'

o
U

20' .,

20

!iulion, .tralght ."".d

2:: f'~
o

20' U

2::l_____

o. J~

20' j
0'

20'

2:: J_~~_ __
20

o
20' ""

20'U~

40' ,iglU

20. J

20' down

20'O~

2' 3~

O'
20'

Fig.11.13. Partial suppression of upbeat nystagmus with baclofen in a patient suffering from a medullary metastasis. Vertical ENG
recordings (top). MR scan and angiography of the vertebral artery (bottom).

Vestibular disorders in (sagittal) pitch plane

Table 11.3 summarises the information given in


this chapter about the upbeat nystagmus/vertigo
syndrome.
Table 11.3. Upbeat nystagmus (vestibular upbeat syndrome)
Clinical syndrome
- Upbeat nystagmus in the primary position of gaze (no suppression by
fixation), modulated by static head tilt
- Associated distressing oscillopsia and postural imbalance
- Saccadic upward pursuit
- Transitions from upbeat to down beat nystagmus possible
Incidence/age/sex
Depends on aetiology, no obvious preference of sex, rare in children
(congenital)
Pathomechanism
Tone imbalance of the vertical semicircular canal reflexes (pitch plane)
modulated by otolithic input
Structural or functionallesions involve:
- either pontomesencephalic junction (brachium conjunctivum? ventral
tegmental tract?)
- or medulla (perihypoglossal nuclei?)
Aetiology
Brainstem tumours, infarction, haematoma, cavernoma, multiple
sclerosis, encephalitis, abscess, alcoholic degeneration (Wernicke's
encephalopathy)' drug intoxication, nicotine (see Table 11.1)
Course/prognosis
Depending on aetiology, gradual improvement by central
compensation?
Usually reversible when caused by intoxication
Management
Medical treatment with baclofen
Physical exercise (eye movements and balance training)
Differential diagnosis
Acquired pendular nystagmus, gaze-evoked nystagmus, down beat
nystagmus, spasmus nutans (infants), vertical congenital nystagmus,
reversed ocular bobbing

References
Abel LA, Traccis S, DeIl'Osso LF, Ansevin CF (1983) Variable waveforms in downbeat nystagmus imply short-term gain changes.
Ann Neurol13:616-620
Alpert JN (1978) Downbeat nystagmus due to anticonvulsant toxicity.Ann NeuroI4:471-463
Baker R, Berthoz A (1975) Is the prepositus hypoglossi nucleus
the source of another vestibulo-ocular pathway? Brain Res
86:121-127
Baker R, Berthoz A (1976) Prepositus neurons synapse directly on
vertical oculomotor neurons. Fed Proc 35:562
Baloh RW, Spooner JW (1981) Downbeat nystagmus: A type of
central vestibular nystagmus. Neurology 31: 304-310
Baloh RW, Yee RD (1989) Spontaneous vertical nystagmus. Rev
Neurol Paris 145:527-532
Barton JJ, Huaman AG, Sharpe JA (1994) Muscarinic antagonists
in the treatment of acquired pendular and downbeat nystagmus: a double-blind study, randomized trial of three intravenous drugs.Ann NeuroI35:319-325

211
Bender MB, Gormen WF (1949) Vertical nystagmus on direct forward gaze with vertical oscillopsia. Am J Ophthalmol
32:967-972
Benjamin EE, Zimmermann CF, Troost BT (1986) Lateropulsion
and upbeat nystagmus on manifestations of central vestibular
dysfunction.Arch NeuroI43:962-964
Bertholon P, Convers P, Barral FG, Duthel R, Michel D (1993)
SyringomytHobulbie posttraumatique et nystagmus vertical
inferieur. Rev NeuroI149:355-358
Bisdorff AR, Wolsley CJ, Anastasopoulus D, Bronstein AM, Gresty
MA (1996) The perception ofbodyverticality (subjective postural vertical) in peripheral and central vestibular disorders.
Brain 119:1523-1534
Bixenman WW (1983) Congenital hereditary downbeat nystagmus. Can J OphthalmoI18:344-348
Blair S, Gavin M (1979) Modification of the macaque's vestibuloocular reflex after ablation of the cerebeIlar vermis. Acta
OtolaryngoI88:225-243
Bogousslavsky J, Regli F (1985) Monocular downbeat nystagmus. J
NeuroI232:99-101
Brandt Th (1991) Vertigo: its multisensory syndromes. Springer,
London
Brandt Th, Dieterich M (1995) Central vestibular disorders in roll,
pitch and yaw planes: topographic diagnosis of brainstem disorders. Neuro-ophthalmology 15:291-303
Brandt Th, Dieterich M, Bchele W (1986) Postural abnormalities
in central vestibular brainstem lesions. In: Bles W, Brandt Th
(eds) Disorders of posture and gait. Elsevier, Amsterdam, pp
141-156
Bronstein AM, Miller DH, Rudge P, Kendall BE (1987)
Downbeating nystagmus: magnetic resonance imaging and
neuro-otological findings. J Neurol Sci 81:173-184
Bchele W, Brandt Th, Degner D (1983) Ataxia and oscillopsia in
downbeat nystagmus/vertigo syndrome. Adv Oto-RhinoLaryngoI30:291-297
Bttner U, Heimchen Ch, Bttner-Ennever JA (1995) The localizing value of nystagmus in brainstem dis orders. Neuroophthalmology 15:283-290
Carl JR, Yee RD, Baloh RW (1982) Convergence and gaze effects on
vertical nystagmus. Invest Ophthal Vision Res (Suppl) 22:265
Carpenter MB, Cowie RJ (1985) Connections and oculomotor projections of the superior vestibular nucleus and ceIl group "y".
Brain Res 336:265-287
Chambers BR, Eil JJ, Gresty MA (1983) Case of downbeat
nystagmus influenced by otolith stimulation. Ann Neurol
13:204-207
Chan T, Logan P, Eustace P (1991) Intermittent downbeat nystagmus secondary to vermian arachnoid cyst with associated
obstructive hydrocephalus. J Clin NeuroophthalmoI11:293-296
Clement G, Vieville'T, Lesteinne F, Berthoz A (1986) Modifications
of gain asymmetry and beating field of vertical optokinetic
nystagmus in microgravity. Neurosci Lett 63:271-274
Cogan DG (1968) Downbeat nystagmus. Arch Ophthalmol
80:757-768
Coppeto JR, Monteiro MLR, LeseIl S, Bear L, Martinez-Maldonado
M (1983) Downbeat nystagmus: Long-term therapy with moderate-dose lithium carbonate. Arch NeuroI40:754-755
Corbett JJ, Jacobsen DM, Thompson HS, Hart MN, Albert DW
(1989) Downbeating nystagmus and other ocular motor defects
caused by lithium toxicity. Neurology 39:481-486
Costin JA, Smith JL, Emery S, Tomsak, RL (1980) Alcoholic downbeat nystagmus.Ann OphthalmoI12:1127-1131
Cox TA, Corbett JJ, Thompson HS, Lennarson L (1981) Upbeat
nystagmus changing to downbeat nystagmus with convergence.
Neurology 31:891-892
Crevits L, Reynaert C (1991) Posture dependent direction reversal
of spontaneous vertical nystagmus. Neuro-ophthalmology
11:285-287

212
Currie J, Matsuo V (1986) The use of clonazepam in the treatment
of nystagmus induced oscillopsia. Ophthalmology 93:924-932
DaroffRB, Troost BT (1973) Upbeat nystagmus. JAMA 225:312
DeJong JMBV, Cohen B, Matsuo V, Uemura T (1980) Midsagittal
ponto-medullary brainstem section: effects on ocular adduction and nystagmus. Exp NeuroI68:420-442
Dichgans J, Mauritz KH, Allum JHJ, Brandt Th (1976) Postural
sway in norm als and atactic patients: analysis of the stabilizing
and destabilizing effects of vision. Aggressologie 17:15-24
Dieterich M, Brandt Th (1987) Impaired motion perception in
congenital nystagmus and acquired ocular motor palsy. Clin
Vision Sci 1:337-345
Dieterich M, Straube A, Brandt T, Paulus W, Bttner U (1991) The
effects of baclofen and cholinergic drugs on upbeat and downbeat nystagmus. J Neurol Neurosurg Psychiatry 54:627-632
Dieterich M, Grnbauer WM, Brandt Th (1998) Direction-specific
impairment of motion perception and spatial orientation in
downbeat and upbeat nystagmus in humans. Neurosci Lett
245:29-32
Fisher A, Gresty M, Chambers B, Rudge P (1983) Primary position
upbeating nystagmus: a variety of central positional nystagmus. Brain 106:949-964
Forsythe WI (1955) Congenital hereditary vertical nystagmus. J
Neurol Neurosurg Psychiatry 18:196-198
Fujikane M, Katayama S, Hirata K, Sunami S (1992) Central diabetes insipidus complicated with upbeat nystagmus and
cerebellar ataxia. Rinsho-Shinkeigaku 32:68-70
Furman JM, Baloh RW, Yee RD (1986) Eye movement abnormalities in a family with cerebellar vermian atrophy. Acta
Otolaryngol (Stockh) 101:371-377
Gilman N, Baloh RW, Tomiyasu U (1977) Primary position upbeat
nystagmus. Neurology 27:294-298
Gresty MA, Baratt H, Rudge P, Page N (1986) Analysis of downbeat
nystagmus. Otolithic versus semicircular canal influences. Arch
NeuroI43:52-55
Gresty MA, Bronstein AM, Brookes GB, Rudge P (1988) Primary
position upbeating nystagmus associated with middle ear disease. Neuro-ophthalmology 8:321-328
Haddad GM, Demer JL, Robinson DA (1980) The effect oflesions
of the dorsal cap of the inferior olive on the vestibulo-ocular
and optokinetic system of the cat. Brain Res 185:265-275
Halmagyi GM, Rudge P, Gresty MA, Leigh RJ, Zee DS (1980)
Treatment of periodic alternating nystagmus. Ann Neurol
8:609-611
Halmagyi M, Rudge P, Gresty MA, Sanders MD (1983)
Downbeating nystagmus, a review of 62 cases. Arch Neurol
40:777-784
Halmagyi GM, Lessell I, Curthoys IS, Lessell S, Hoyt WF (1989)
Lithium-induced downbeat nystagmus. Am J Ophthalmol
107:664-670
Hammack H, Kotanides H, Rosenblum MK, Posner JB (1992)
Paraneoplastic cerebellar degeneration. 11. Clinical and
immunologic findings in 21 patients with Hodgkin's disease.
Neurology 42:1983-1943
Harada K, Kato I, Nakamura T, Koike Y (1994) Role of the prepositus hypoglossi nucleus on primary position upbeat nystagmus.
Acta Otolaryngol (Stockh) SuppI511:120-125
Hiel H, Elgoyhen AB, Drescher DG, Morley BJ (1996) Expression
of nicotinic acetylcholine receptor mRNA in the adult rat
peripheral vestibular system. Brain Res 738:347-352
Hirose G, Kawada J, Tsukada K, Komatsuzaki A, Sharpe JA (1991)
Primary position upbeat nystagmus. Acta Otolaryngol (Stockh)
SuppI481:357-360
Hoyt CS, Gelbert SS(1984) Vertical nystagmus in infants with congenital ocular abnormalities. Ophthal Paediatr Genet
(Amsterdam) 4:155-162
Hwang TL, Still CN, Jones JE (1995) Reversible downbeat nystagmus and ataxia in febamate intoxication. Neurology 45:846

Vertigo
Igarashi M, Takahashi M, KuboT, Levy JK, Homic JL (1978) Effect
of macular ablation on vertical optokinetic nystagmus in the
squirrel monkey. ORL 40:312-318
Ishiyama A, Lopez I, Wackym PA (1995) Distribution of efferent
cholinergic terminals and alpha-bungarotoxin binding to putative nicotinic acetylcholine receptors in the human vestibular
end-organs. Laryngoscope 105:1167-1172
Ho M, Nisimaru N, Yamamoto M (1976) Pathways for the vestibuloocular reflex excitation arising from semicircular canals of
rabbits. Exp Brain Res 24:257-271
Iwata A, Takao F, Kunimoto M, Inoue K (1996) Primary position
upbeat nystagmus reversed with carbamazepine. Eur J Neurol
3:260-263
Jacobson DM, Corbett JJ (1989) Downbeat nystagmus associated
with dolichoectasia of the vertebrobasilar artery. Arch Neurol
46: 1005-1008
Jacome DE (1986) Monocular downbeat nystagmus. Ann
OphthalmoI18:293-296
Kato I, Nakamura T, Watanabe J, Harada K, Aoyagi M, Katagiri T
(1985) Primary position upbeat nystagmus, localizing value.
Arch Neurol42: 819-821
Kattah JC, Dagi TF (1990) Compensatory head tilt in upbeating
nystagmus.J Clin Neuro-ophthalmollO:27-31
Keane JR, Itabashi HH (I987) Upbeat nystagmus: clinicopathologic
study of two patients. Neurology 37:491-494
Lavin PJM, Traccis S, Dell'Osso LF, Abel LA, Ellenberger C jr
(1983) Downbeat nystagmus with a pseudocycloid waveform:
Improvement with base-out prisms. Ann NeuroI13:621-624
Lawden MC, Bronstein AM, Kennard C (1995) Repetitive paroxysmal nystagmus and vertigo. Neurology 45:276-280
Leigh RJ, Zee DS (1991) The neurology of eye movements. 2nd Ed.
Davis, Philadelphia
Malm G, Lyiug- Tunell U (1980) Cerebellar dysfunction related to
toluene sniffing.Acta Neurol Scand 62:188-190
Matsuo V, Cohen B (1984) Vertical optokinetic nystagmus and
vestibular nystagmus in the monkey: Up-down asymmetry and
effects of gravity. Exp Brain Res 53:197-216
Mayfrank L, Thoden U (1986) Downbeat nystagmus indicates
cerebellar or brain-stem lesions in vitamin B12 deficiency. J
Neurol 233: 145-148
McConnell HW, Darlington CL, Smith PF, Sturge DL, Thomson SD,
Nukada H, Mair MW (1990) Hereditary cerebellar degeneration
with downbeat nystagmus. A case and its treatment. Acta
Neurol Scand 81:423-426
Mizuno M, Kudo Y, Yamane M (1990) Upbeat nystagmus influenced by posture: report of two cases. Auris-Nasus-Larynx
16:215-221
Mochizuki Y, Oishi M, Hagi C, Iida S (1996) Ipsilateral oculomotor
nerve palsy and contralateral downbeat nystagmus due to unilateral paramedian thalamopenduncular infarction. A case
report. Clin NeuroI36:800-802
Monteiro ML, Sampaio CM (1993) Lithium-induced downbeat
nystagmus in a patient with Arnold-Chiari malformation. Am J
OphthalmoII16:648-649
Munro NAR, Gaymard B, Rivaud S, Majdalani A, PierrotDeseilligny Ch (1993) Upbeat nystagmus in a patient with a
small medullary infarct. J Neurol Neurosurg Psychiatry
56:1126-1128
Murphy MG, O'Leary JL (1971) Neurological deficit in cats with
lesions of the olivocerebellar system. Arch Neurol 24: 145-157
Nakada T, Remler MP (1981) Primary position upbeat nystagmus.
J Clin Neuroophthalmol 1:185-189
Neveling R, Kruse KE (1961) ber Nicotinnystagmus. Arch Ohr
Heilk Z Hals Heilk 177:427-431
O'Brian FH, Bender MB (1945) Localizing value ofvertical nystagmus. Arch Neurol Psychiatry 54:378-380
Oliva A, Rosenberg ML (1990) Convergence-evoked nystagmus.
Neurology 40:161-162

Vestibular disorders in (sagittal) pitch plane


Pawl R (1973) Upbeat nystagmus. JAMA 225:312
Pedersen RA, Troost BT, Abel LA, Zomb D (1980) Intermittent
downbeat nystagmus and oscillopsia reversed by suboccipital
craniectomy. Neurology 30:1239-1242
Phelan KD, Gallagher JP (1992) Direct muscarinic and nicotinic
receptor-mediated excitation of rat medial vestibular nucleus
neurons in vitro. Synapse 10:349-358
Pine! JF, Larmande P, Guegan Y, Iba-Zizen MTh (1987) Downbeat
nystagmus: Case report with magnetic resonance imaging and
surgical treatment. Neurosurgery 21: 736-739
Precht W, Anderson WH, Blanks RH! (1979) Canal-otolith convergence on cat ocular motorneurons. In: Granit R, Pompeiano 0
(eds) Reflex control of posture and movement. Progress in
brain research, vol 50. Elsevier, Amsterdam, pp 459-468
Ranalli PI, Sharpe JA (1988) Upbeat nystagmus and the ventral
tegmental pathway of the upward vestibulo-ocular reflex.
Neurology 38: 1329-1330
Rosengart A, Hedges TR, Teal PA, DeWitt LD, Wu JK, Wolpert S,
Caplan LR (1993) Intermittent downbeat nystagmus due to vertebral artery compression. Neurology 43:216-218
Rousseaux M, Dupard T, Lesoin F, Barbaste P, Hache JC (1991)
Upbeat and downbeat nystagmus occurring successively in a
patient with posterior medullary haemorrhage. J Neurol
Neurosurg Psychiatry 54:367-369
Rude SA, Baker JF (1996) Otolith orientation and downbeat
nystagmus in the normal cat. Exp Brain Res 111:144-148
Sakuma A, Kato I, Ogino S, Okada T, Takeyama 1(1996) Primary
position upbeat nystagmus with special reference to alteration
to downbeat nystagmus. Acta Otolaryngol (Stockh) Suppl
522:43-46
Saul RF, Selhorst JB (1981) Downbeat nystagmus with magnesium
depletion. Arch Neurol 38: 650-652
Schmidt D (1972) ber die Bedeutung der Augensymptome zur
Diagnostik der Wernicke Encephalopathie. Klin Mbl
Augenheilk 161:36-42
Senelick RC (1981) Total alleviation of downbeat nystagmus in
basilar impression by transoral removal of the odontoid
process. J Clin Neuroophthalmoll:265-267
Shibasaki H, Yamashita Y, Motomura S (1978) Suppression of congenital nystagmus. J Neurol Neurosurg Psychiatry 41:1078-1083
Sibony, PA, Evinger C, Manning KA (1987) Tobacco-induced
primary position upbeat nystagmus. Ann Neurol 21:53-58
Sibony PA, Evinger C, Manning KA, Pellegrini JJ (1988) The effects
of tobacco smoking on smooth pursuit eye movements. Ann
NeuroI23:238-241
Sibony PA, Evinger C, Manning KA (1990) Nicotine and tobaccoinduced nystagmus. Ann Neurol28: 198

213
Sogg RL, Hoyt WF (1962) Intermittent vertical nystagmus in a
father and son. Arch Ophthal (Chicago) 68:515-517
Spiegel EA, Scala NP (1941) Vertical nystagmus following lesions
of the cerebellar vermis. Arch Ophthalmol 26:661-669
Spooner JW, Baloh RW (1981) Arnold-Chiari malformation:
Improvement in eye movements after surgical treatment. Brain
104:51-60
Stengel E (1935) Zur Frage der Herdlokalisation bei spontanem
Vertikalnystagmus. Z Ges Neurol Psychiatr 153:417-424
Sullivan JD jr, Rumack BH, Peterson RG (1981) Acute carbamazepine toxicity resulting from overdose. Neurology
31:621-624
Takahashi M, Sakurai S, Kanzaki J (1987) Horizontal and vertical
optokinetic nystagmus in man. ORL 40:43-53
Takemori T, Suzuki M (1977) Cerebellar contribution to oculomotor
function. Oto-Rhino-Laryngol 39:209-217
Tokumasu K, Fujino A, Yoshio S, Nitta K, Goto K, Yoneda S (1991)
Upbeat nystagmus in primary eye position. Acta Otolaryngol
(Stockh) SuppI481:366-368
Trobe JD, Sharpe JA, Hirsch DK, Gebarski SS (1991) Nystagmus of
Pelizaeus-Merzbacher disease. A magnetic search-coil study.
Arch NeuroI48:87-91
Troost BT, Martinez J, Abel LA, Heros RC (1980) Upbeat nystagmus and internuclear ophthalmoplegia with brainstem glioma.
Arch NeuroI37:453-456
Weissman BM, Dell'Osso LF, Discenna A (1988) Downbeat nystagmus in an infant: spontaneous resolution during infancy.
Neuro-ophthalmology 8:317 -319
Wheeler SD, Ramsay RE, Weiss J (1982) Drug-induced downbeat
nystagmus.Ann NeuroI12:227-228
Yamazaki K, Katayama S, Ishihara T, Hirata K (1994) A case of
Fisher's syndrome with upbeat nystagmus. Rinsho-Shinkeigaku
34:489-492
Yee RD, Baloh RW, Honrubia V (1983) Episodic vertical oscillopsia
and downbeat nystagmus in a Chiari malformation. Arch
OphthalmoI102:723-725
Zasorin NL, Baloh RW (1984) Downbeat nystagmus with alcoholic
cerebellar degeneration. Arch Neurol 41: 130 1-1302
Zee DS, Friendlich AR, Robinson DA (1974) The mechanism of
downbeat nystagmus. Arch Neurol 30:227 - 23 7
Zee DS, Yamazaki A, Butler PH, Gucer G (1981) Effects of ablation
of flocculus and paraflocculus on eye movements in primate. J
NeurophysioI46:878-899
Zumstein HR, Meienberg 0 (1982) Upbeat nystagmus and visual
system disorder in Wernicke's encephalopathy due to starvation. Neuro-ophthalmology 2: 157 -162

Vestibular disorders in (horizontal)


yaw plane

or pitch (p. 170) covers nearly the entire brainstem from the medulla to the rostral midbrain
(Figs C.2 and 10.3). The larger extent of the latter
area is due to the greater separation of the
vestibular nuclei and the ocular motor integration centers for vertical and torsional eye movements (riMLF and INC).
Second, the area of alesion that can theoretically
cause a pure tone imbalance in the yaw plane
adjoins and overlaps areas subserving vestibular
function in roll and pitch (Figs C.2 and C.3).
There is a multisensory convergence within the
parallel neural network of the vestibular nuclei
(Leigh and Brandt 1993), alesion of which will
cause mixed vestibular syndromes in more than
one plane. A study of vestibular nuclei lesion in
the monkey demonstrated a combined nystagmus: its horizontal component beat toward the
contralateral side after rostraliesions and toward
the ipsilateral side after caudallesions (Uemura
and Cohen 1973).

The clinical signs, both perceptual and motor, of a


vestibular tone imbalance in the yaw plane include
rotational vertigo, deviation of perceived straightahead, pastpointing, rotational and lateral body faUs
and horizontal nystagmus.
--------------------~

Horizontal nystagmus as a sign of


vestibular tone imbalance in the yaw
plane
Spontaneous nystagmus is present when the eye is at
midposition in the orbit, i.e. during straight-ahead
gaze. This eye position usually differs from the socalled primary position of the eyes, a term that refers
to the origin of the coordinate system for Listing's
plane (Bttner et a1. 1995). Spontaneous nystagmus
is usually considered a sign of peripheral or central
vestibular tone imbalance. This imbalance can be
caused by unequal activity patterns of bilateral
vestibular pathways or by a neural integrator defect.
Whereas the first is characterised by constant velocity
slow phases, the neural integrator deficit is characterised by exponentially decaying slow phases.
Furthermore, an imbalance of the smooth pursuit
system or defects of the saccade generators may
cause nystagmus. Central vestibular syndromes
manifesting purely in the yaw plane occur less
frequently than those due to imbalance in the vertical pitch and roll planes for two reasons (Brandt and
Dieterich 1995).

First, the area of alesion that can cause a tone


imbalance in yaw is comparatively small (root
entry zone of the vestibular nerve, medial and
superior vestibular nuclei, and the adjacent
integration cent re for horizontal eye movements,
the paramedian pontine reticular formation
(PPRF. In contrast, the area of alesion that can
cause vestibular tone imbalance in roll (p. 179)

Horizontal benign paroxysmal positioning vertigo


(McClure 1989; Pagnini et a1. 1989; Baloh et a1. 1993;
p. 269) is one of the few vestibular disorders restricted
to the yaw plane. Here the positional nystagmus is
linear-horizontal due to the stimulation of one horizontal semicircular canal only.
Some of the cases described as central variants of
vestibular neuritis (Kmpf 1986; Hopf 1987;
Dieterich and Bchele 1989; Disher et a1. 1991;
Francis et a1. 1992) and caused by lesions of the
medial vestibular subnucleus or the root entry zone
of the vestibular nerve were probably not restricted
to the yaw plane, because the case descriptions also
contain signs and symptoms ofVOR tone imbalance
in other planes of action. There have been frequent
reports that cerebellar infarctions due to occlusion
of the anterior inferior cerebellar artery (which also
supplies the rostral vestibular nuclei) mimic
vestibular neuritis (Kmpf 1986; Amarenco et a1.
1990). The other main cause of confusion with disorders at the entry zone of the eighth nerve is multiple

215

Vertigo

216

sclerosis (Brandt et al. 1986; Dieterich and Bchele


1989).
Hopf (1987) has described two patients with
symptoms and signs of vestibular neuritis, including
horizontal semicircular canal paresis, who had in
addition paresis of the masseter, temporal, and pterygoid muscles as weH as impairment of the masseter
reflex on the affected side. He argued that infarction
of a smaH region ventrolateral to the floor of the
fourth ventricle was causative, as structures serving
both peripheral vestibular function and trigeminal
motor function are present in this area. A demyelinating plaque in multiple sclerosis may mimic signs
of vestibular neuritis, including caloric hyporesponsiveness if it is located at the root entry zone of the

eighth nerve (Fig. 12.1; Brandt et al. 1986). It is


extremely important to investigate carefuHy oculomotor function in patients presenting with vestibular
dysfunction which is apparently of peripheral origin,
since most central cases have some abnormalities
incompatible with a peripheral disorder such as a
contralateral gaze-evoked nystagmus or horizontal or
vertical saccadic pursuit. The degree of horizontal
canal paresis (as tested by caloric irrigation) is usually less severe (incomplete) than in peripheral canal
paresis. An MRI study demonstrated that canal paresis caused by brainstem lesions predominantly
involved the medial and to a lesser extent the lateral
vestibular nucleus as weH as the proximal portion of
the vestibular fascicle (Francis et al. 1992).

P.R., d, 46y
spontaneous nystagmus (horizontal ENG)

44 C

right ear

left ear

lett ear

5s
Fig.12.1. Magnetic resonance tomography in a patient (aged 46 years) with an MS plaque of the pontomedullary tegmentum probably involving the entry zone of the left eighth nerve (top). This plaque mimicked a peripherallabyrinthine lesion with contraversive
rotational spontaneous nystagmus, ipsiversive direction of fall, nausea and emesis as weil as an ipsilateral hyporesponsiveness of the
horizontal semicircular canal to caloric irrigation. Later the direction of the spontaneous nystagmus was reversed (bottom); caloric irrigation still revealed hyporesponsiveness of the left horizontal semicircular canal. (From Brandt et al. 1986.)

217

Vestibular disorders in (horizontal) yaw plane

Combined VOR dysfunction in more


than one plane of action
Vestibular syndromes - when caused by unilateral
pontomedullary lesions - frequently result in combined vestibular tone imbalance in more than one
plane, such as a combination of torsional and horizontal nystagmus. This tone imbalance may manifest
not only in spontaneous nystagmus but also in spontaneous or gaze-evoked ocular deviations. There
may be an inappropriate horizontal ocular deviation
during attempted vertical saccades (lateropulsion in
Wallenberg's syndrome) or an inappropriate
torsional deviation during attempted horizontal
saccades ("torsipulsion"; Morrow and Sharpe 1988;
FitzGibbon et al. 1996).
Sometimes the clinical manifestation of a particular ocular motor abnormality such as ocular tilt
reaction allows one to identify the semicircular canal
pathway affected (anterior or posterior). The different presentation of ocular tilt reaction in
Wallenberg's syndrome (p. 309) with monocular dissociation of ocular torsion - either excyclotropia of
the undermost eye or incydotropia of the uppermost eye - indicates involvement of the posterior or
the anterior semicircular canal pathways (Dieterich
and Brandt 1992). Lesions of the vestibular nudei
mayaiso result in repetitive multidirectional paroxysmal nystagmus and vertigo, which have been
reported to respond to treatment with carbamazepine (Lawden et al. 1995). The differential
effects of a medullary lesion involving more than
one VOR plane are not only reflected in the direction
of eye movements but also in the preferred direction
of increased body sway. Body sway histograms as
measured by posturography are primarily diagonal
in patients with Wallenberg's syndrome and moderate body lateropulsion (combination: roll and yaw),
but primarily lateral in patients with severe body
lateropulsion (roll> yaw) (Fig. 19.3; Dieterich and
Brandt 1992). This is also reflected by the dose correlation between the tilt of perceived visual vertical
and the severity of body lateropulsion, both indicators - perceptual and postural - of a vestibular
imbalance in roll (Fig. 19.4; Brandt and Dieterich
1993).
. It is sometimes emphasised that vestibular neuritis is a prototype of a tone imbalance in yaw plane.
However, the obligatory sign of a combined horizontal!
torsional nystagmus contraversive to the lesion side
unequivocally indicates the combined impairment of
inputs from the horizontal (vertical) and anterior
semicircular canals (Bchele and Brandt 1988; Fetter
and Dichgans 1996) and the otoliths. Thus, vestibu-

lar neuntIs combines dis orders in yaw and roll


planes: the nystagmus is secondary to horizontal and
anterior semicircular canal dysfunction, and the
tonic signs (ocular torsion, tilt of perceived vertical)
are secondary to otolith dysfunction. Tonic torsion
of both eyes and tilts of the subjective visual vertical
are typical features of an acute vestibular neuritis
(Bhmer and Rickenmann 1995) or vestibular nerve
neurectomy (Curthoys et al. 1991).
Caution should be exercised when trying to identify functional loss of one or more semicircular
canals by simply analysing the directions of spontaneous vestibular nystagmus. Bhmer et al. (1997)
found that 3-D analysis of angular velo city vectors of
the spontaneous nystagmus induced by vestibular
neurectomy or by vestibular neuritis could not distinguish them; both dustered along the sensitivity
vector of the lateral semicircular canal. These authors
conduded that "the absence of nystagmus components in directions of the vertical semicircular
canals may reflect an anisotropy of the oculomotor
efferent part of the VOR arc rather than lesions limited to the lateral semicircular canal afferents. The 3D analysis of spontaneous nystagmus therefore does
not per mit accurate localization of a peripheral
vestibular lesion". Analysis of dynamic 3-D ocular
responses to angular acceleration in the plane of the
affected canal may prove to be more specific.

References
Amarenco P, Roullet E, Chemouilli Ph, Marteau R (1990) Infaretus
pontin inferolateral: deux aspeets cliniques. Rev Neurol (Paris)
146:433-437

Baloh RW, Jaeobson K, Hornrubia V (1993) Horizontal semicireu


lar eanal variant of benign position al vertigo. Neurology
43:2542-2549

Bhmer A, Rickenmann J (1995) The subjective visual vertical as a


clinieal parameter of vestibular funetion in peripheral vestibu
lar diseases. J Vestib Res 5:35-45
Bhmer A, Straumann D, Fetter M (1997) Threedimensional
analysis of spontaneous nystagmus in peripheral vestibular
lesions. Ann Otol-Rhinol-Laryngol 106:61-68
Brandt Th, Dieterich M (1993) Vestibular falls. J Vestib Res 3:3-14
Brandt Th, Dieterich M (1995) Central vestibular syndrome in
roll, pitch, and yaw planes. Topographie diagnosis of brainstem
dis orders. Neuro-ophthalmology 15:291-303
Brandt Th, Dieterich M, Bchele W (1986) Postural abnormalities
in central vestibular brainstem lesions. In: Bles W, Brandt Th
(eds) Disorders of posture and gait. Elsevier, Amsterdam, pp
141-156

Behele W, Brandt Th (1988) Vestibular neuritis - a horizontal


semicircular canal paresis? Adv Oto-Rhino-Laryngol
42:157-161

Bttner U, HeImchen C, Bttner-Ennever JA (1995) The localizing


value of nystagmus in brainstem disorders. Neuro-opthalmololgy
15:283-290

Curthoys IS, Dai MI, Halmagyi GM (1991) Human ocular torsion al

218
position before and after unilateral vestibular neurectomy. Exp
Brain Res 85:218-225
Dieterich M, Brandt Th (1992) Wallenberg's syndrome: lateropulsion, cyclorotation and subjective visual vertical in 36 patients.
Ann NeuroI31:399-408
Dieterich M, Bchele W (1989) MRI findings in lesions at the
entry zone of the eighth nerve. Acta Otolaryngol (Stockh) Suppl
468:385-389
Disher MJ, Telian SA, Kemink JL (1991) Evaluation of acute vertigo: unusual lesions imitating vestibular neuritis. Am J Otol
12:227-231
Fetter M, Dichgans J (1996) Vestibular neuritis spares the inferior
division of the vestibular nerve. Brain 119:755-763
FitzGibbon EJ, Calvert PC, Dieterich M, Brandt Th, Zee DS (1996)
Torsional nystagmus during vertical pursuit. Neuroophthalmology 16:79-90
Francis DA, Bronstein AM, Rudge P, du Boulay EPGH (1992) The
site of brainstem lesions causing semicircular canal paresis: an
MRI study. J Neurol Neurosurg Psychiatry 55:446-449

Vertigo
Hopf HC (1987) Vertigo and masseter paresis. A new brainstem
syndrome. J NeuroI235:42-45
Kmpf D (1986) Der benigne pseudovestibulre Kleinhirninsult.
Nervenarzt 57:163-166
Lawden MC, Bronstein AM, Kennard C (1995) Repetitive paroxysmal nystagmus and vertigo. Neurology 45:276-280
Leigh RJ, Brandt Th (1993) Are-evaluation of the vestibulo-ocular
reflex: new ideas of its purpose, properties, neural substrate,
and dis orders. Neurology 43:1288-1295
McClure JA (1989) Horizontal canal BPV. Otolaryngology
14:30-35
Morrow MI, Sharpe A (1988) Torsional nystagmus in the lateral
medullary syndrome. Ann NeuroI24:390-398
Pagnini P, Nuti D, Vannuncchi P (1989) Benign paroxysmal
vertigo of the horizontal cana!. ORL-J Otorhinolaryngol Rel
Spec 51:161-170
Uemura T, Cohen B (1973) Effects ofvestibular nuclei lesions on
vestibulo-ocular reflexes and posture in monkeys. ActaOtolaryngol (Stockh) SuppI315:1-71

Vestibular cortex: its locations,


functions, and disorders

The two major cortical functions of the vestibular


system are spatial orientation and self-motion perception. These functions, however, are not exclusively
vestibular; they also rely on visual and somatosensory input. All three systems (vestibular, visual and
somatosensory) provide us with redundant information about the position and motion of our body relative to the external space. Although the vestibular
cortex function is distributed among several multisensory areas in the parietal and temporal cortices, it
is also integrated in a larger network for spatial attention and sensorimotor control of eye and body
motion in space.
In this section on central vestibular disorders
(Section CL the analysis of vestibular cortex function must concentrate on how the system successfully integrates multisensory input into a unique
perception and how it uses this information for
accurate motor performance in space. The analysis
of peripheral vestibular function and disorders
(Section B) concentrates more on how labyrinthine
receptors transduce gravitational force and head
acceleration. Cortical representation is bilateral with
right hemispheric dominance. Neurological syndromes with cortical vertigo include paroxysmal
excitation (vestibular epilepsy, p. 233; room-tilt illusion, p. 224) or lesional dysfunction (tHt of perceived
vertical with lateropulsion, p. 192; spatial hemineglect, p. 224).

(dkvist et a1. 1974), the parieto-insular vestibular


cortex (PIVC) at the posterior end of the insula
(Grsser et a1. 1990a,b, 1995), and area 7 in the
inferior parietal lobule (Faugier-Grimaud and
Ventre 1989) (Fig. 13.1). Not only do these areas
receive bilateral vestibular input from the vestibular
nu dei, but they in turn direct1y project down to the
vestibular nudei (Abkarian et a1. 1994; Guldin and
Grsser 1996; Jeannerod 1996). Thus, corticofugal
feedback may modulate vestibular brainstem
function.
Our knowledge about vestibular cortex function
in humans is less precise. It is derived mainly from
stimulation experiments reported anecdota11y in
the older literature and from recent brain activation
studies with PET (Bottini et a1. 1994) and fMRI
(Dieterich et a1. 1998a; Bucher et a1. 1997, 1998;
Brandt et a1. 1998a). It is not always possible to
extrapolate from monkey species to human cortex,
as Andersen and Gnadt (1989) demonstrated for
Brodmann's area 7 in rhesus monkey and humans.
Area 2v corresponds best to the vestibular cortex as
described by Foerster (1936). The PIVC corresponds best to a region from which Penfield and
Jasper (1954) were able to induce vestibular sensations by electrical stimulation with a depth electrode within the sylvian fissure, medial to the
primary acoustic cortex (compare Fig. 14.1).

No primary vestibular cortex

Multiple vestibular cortex areas


Animal studies have identified several distinct and
separate areas of the parietal and temporal cortices
which receive vestibular afferents, such as area 2v at
the tip of the intraparietal sulcus (Schwarz and
Fredrickson 1971; Fredrickson et a1. 1966; Bttner
and Buettner 1978), area 3aV (neck, trunk and
vestibular region of area 3a) in the central SUlcUS

The very multi pli city of the representations of


vestibular cortical areas raises doubts about the
existence of a primary vestibular cortex comparable
to the visual or the auditory cortex. Microelectrode
recordings from the so-ca11ed vestibular areas a11
demonstrate that the neurons are multisensory:
they respond not only to vestibular but also to
somatosensory and optokinetic stimuli (PIVe:
Grsser et a1. 1990a,b; area 2v: Bttner and Buettner

219

220

Vertigo

3a
7 a, b

PIVC

3a

Fig.13.1. Schematic representation of a monkey brain (top)


with the experimentally established areas that receive vestibular
input: area 2v at the anterior part of the intraparietal sulcus, area
3a in the central sulcus, multisensory area 7 at the inferior parietal cortex, and the parieto-insular vestibular cortex (PIVC) deep
in the posterior end of the insula. The schematic representation
of the human brain (bottam) indicates the postulated homologue ofthe monkey areas. c = central sulcus; ip = intraparietal
sulcus; I = lateral sylvian sulcus; ts = superior temporal sulcus.
(From Brandt et al. 1995.)

1978; area 3a: dkvist et al. 1974, Phillips et al. 1971,


Schwarz et al. 1973). The same is true for area 7 of
the inferior parietal cortex (Faugier-Grimaud and
Ventre 1989; Andersen 1987; Kawano et al. 1980), a
major multisensory integration centre for spatial
orientation and visuomotor function. Area 7 also
receives vestibular input, and vestibular neurons
accumulated in part of area 7 (7ant) may represent
the homologue to area 2v of the macaque brain
(Guldin et al. 1992). The cytoarchitectonic structure
of these vestibular cortical areas is also more characteristic of a multisensory or sensorimotor rather
than a primary (unimodal) sensory cortex. Pandya
and Sanides (1973) emphasised the cytoarchitectonic homology of the retroinsular parietal cortex
(reipt) in the monkey and the vestibular area anterior to the suprasylvian sulcus (ASSS) in the cat

(Walzl and Mountcastle 1949; Mickle and Ades


1952).
If separate primary cortices for visual and auditory signals were established by evolution only to be
abandoned later to provide the vestibular system
with a primary area, what effect would this have on
the conscious perception of different sensory
modalities? Why is there no purely unimodal
vestibular sensation? Shape and colour of a presented object are detected by vision alone, and analysis
within the visual cortex does not require proprioception or other senses. The same is true for the differentiation of tones and melodies. In contrast,
natural stimulation of the vestibular system during
head motion and locomotion is always multisensory (visual, vestibular, somatosensory). Redundant
information for spatial orientation and postural
control is provided by different sensory cues,
among which vestibular signals may playa dominant role. Clinical vertigo syndromes with disorientation and falls serve as an example. Due to
multisensory interaction, vestibular stimulation
causes sensations that have vestibular, somatosensory and visual qualities. Unlike the complexity of
visual and auditory stimuli, the physical characteristics of vestibular stimuli are sufficiently defined
by direction and amount of acceleration applied to
the head. Their concurrent consequences for the
visual and somatosensory systems make aseparate
evaluation of the unimodal vestibular quality in the
cortex unnecessary.

The parieto-insular vestibular cortex (PIVC)


In view of the strong interconnections between PIVC
and other vestibular cortex areas (mainly 3a V and
2v) as weIl as the vestibular brainstem nuclei, Guldin
and Grsser (1996) postulate that it is the core
region within the vestibular cortical system. About
50% of the neurons in this region respond to
vestibular stimulation in addition to somatosensory,
optokinetic, or visual stimulation. This area is
involved not only in the processing of vestibular,
somatosensory and visual information, which is generated whenever the position of the body changes in
relation to the extrapersonal space (Guldin and
Grsser 1996), but also when stationary human subjects perform optokinetic nystagmus (Bucher et al.
1997; Dieterich et al. 1998a).
The terminology and parcelling of the insula and
its surrounding opercula according to cytoarchitectonic structure vary for different species. The overlapping areas of infarctions in patients with

Vestibular cortex: its locations, functions, and disorders

pathological tilts of the subjective vertical centre on


the posterior insula (Chap. 19, Fig. 19.13), a region
that Pandya and Sanides (1973) termed "reipt"
(retroinsular parietal cortex) in the rhesus monkey
and which Grsser and co-workers (1990a,b) delineated as PIVC, apart of the retroinsular cortex (Ri).
The PIVC probably extends into the cervical representation of the secondary somatosensory area
(SII) of the parietal operculum (Jones and Burton
1976). The neighbouring area caudal to the PIVC is
a secondary auditory area. This corresponds with
findings of stimulation experiments in humans
(Penfield and Jasper 1954) and clinical experience
(Smith 1960) that shows that tinnitus and contralateral paraesthesia may precede or accompany vertigo in vestibular epilepsy. Penfield and Kristiansen
(1951) localised the "epileptogenic lesions" in
patients with vestibular au ras to be roughly in the
centre of the posterior part of the temporal lobe. A
focal activation of regional cerebral blood flow
(rCBF) in the superior temporal region posterior to
the auditory area was found during caloric vestibular stimulation in humans (Friberg et al. 1985). The
same areas in the posterior insula, and the temporoparietal cortex, the putamen and the anterior
cingulate gyrus were activated in PET studies during caloric vestibular stimulation in humans
(Bottini et al. 1994; Dieterich et al. 1996).
Surprisingly the posterior insula (PIVC) was deactivated during large-field visual motion stimulation
which induces apparent self-motion ("reciprocal
inhibitory visual-vestibular interaction", see p. 225)
(Brandt et al. 1998a). Furthermore, PIVC is activated in fMRI with a significantly right-hemispheric
dominance, when optokinetic nystagmus is performed (Fig. 13.2). This is not seen if optokinetic
nystagmus is suppressed by fixation (Dieterich et al.
1998a).
Speculations about the existence of a vestibular
cortex function are based on the multiple projections from the thalamus to a kind of"inner circle" of
the vestibular cortical representations, mainly areas
PIVC (Ri, reipt), 7ant (2v) and 3av, which have been
analysed by intracortical retrograde tracer injections
(Akbarian et al. 1992). The "proprioceptive vestibular area 3av" receives its major thalamic projection
from the oral and the superior ventroposterior nucleus (VPo). As the dominant cortical vestibular area,
the PIVC receives its main input from the vestibular
parts of the ventroposterior complex and the medial
pulvinar. Area 7ant receives its major input from the
posteromedial pulvinar. Cortico-cortical connections involve frontal and several parietal cortical
areas as weIl as the parietotemporal association area
T3, which receives optokinetic information via the

221

medial, lateral and inferior pulvinar (Akbarian et al.


1992).

Multimodal sensorimotor vestibular


cortex function and dysfunction
The vestibular, the visual and the somatosensory
systems cooperate to determine our internal representation of space and subjective body orientation in
unique 3-D coordinates, wh ich are either egocentric
(body-centred) or exocentric (world-centred). This
is not a trivial process, since two of the sensory
systems are anchored in the head, which moves relative to the trunk. Retinal coordinates - dependent as
they are on gaze and head position - and head-fixed
labyrinthine coordinates would require continuous
updating of the particular eye and head positions in
order to deliver reliable input for adequate ocular
motor and motor exploration of space. Nature seems
to have solved this impossible sensorimotor control
of a multilink and multiaxis system by multisensory
co ding of space in either common egocentric or exocentric rather than retinotopic or head-centred coordinates. This has been demonstrated for posterior
parietal neurons (Andersen et al. 1985; GaIletti et al.
1993). Spatial information in non-retinal coordinates
allows us to determine body position relative to
visual space, which is a necessary prerequisite for
accurate motor response. To obtain such a frame of
reference, information coded in coordinates of the
peripheral sensory organs (retina, otoliths, semicircular canals and proprioceptors such as muscle
spindies) must be transformed and integrated
(Karnath 1994). This function is most probably subserved by the posterior parietal cortex, alesion of
which pro duces a visuospatial hemineglect. Karnath
et al. (1991; 1993) argued that neglect in braindamaged patients is caused by a disturbance of the
central transformation process that converts the
sensory input coordinates from the periphery into
an egocentric, body-centred co ordinate system
(p. 224). The importance of the vestibular input for
spatial orientation and the continuous updating of
our internal representation of space becomes evident in light of the deficient spatial memory in
microgravity during spacecraft missions. Large
errors are made du ring prolonged microgravity
when pointing at memorised targets, and it is the
lack of knowledge of target position, not limb position, that is causative (Watt 1997).
In patients an inappropriate vestibular input due
to peripheral or, central dysfunction can cause

222

Vertigo

a
Fig.13.2. a fMRI during optokinetic stimulation. T2 *-weighted coronal MR images of five cortical sections with superimposed activation map associated with optokinetic nystagmus induced by right (a, c, e, g, i) and left rotation of a drum (b, d, f. h,j) of a 27-year-old
control subject (TRfTE = 63/ 30 ms, IX = 10). The colour-coded correlation coefficient scale ranges from 0.5 to a maximum of 1.0. During
OKN bilaterally activated areas are the medial part of the superior frontal gyrus (supplementary eye field; S; a-b), the prefrontal cortex
(PF; a-f). the precentral and posterior median frontal gyrus (frontal eye fields; F; a-f) and parts of the parietal cortex including the parietal eye field (P; c-j). The lateral occipitotemporal cortex (0; c-j) shows a strong asymmetric activity, mainly in the right hemisphere.

223

Vestibular cortex: its locations, functions, and disorders

Note that there is no difference in the anatomicallocation and extent of activation for both directions of object motion. b Magnified
cortical and subcortical activation maps superimposed on the corresponding coronal T/-weighted anatomical images ofthe lower
five sections of the subject shown in a. The activation maps demonstrate significant bilateral activity in the anterior part (AI; e-j) and
posterior part (PI; g-h) of the insula (parieto-insular vestibular cortex), the putamen (PU; e-h), the globus pallidus (GP; g-jl, the paramedian thalamus (T; e-f) and the primary visual cortex (VC; a-j). (From Sucher et al. 1997.)

Vertigo

224

paroxysmal "room-tiIt illusions" (p. 224), the resuIt


of amismatch of the two 3-D visual and vestibular
co ordinate maps. Furthermore, a plane- and direction-specific tilt of statie spatial orientation occurs in
disorders of the vestibulo-ocular reflex, such as downbeat and upbeat nystagmus {p. 20l}. Adjustments of
subjective straight ahead exhibit an upward shift in
downbeat nystagmus and a downward shift in upbeat
nystagmus (Dieterich et al. 1998b). Here the tiIt of
perceived straight-ahead is elicited by the asymmetrie
vestibular tone in the pitch plane in the brainstem
which reaches the cortex by ascending projections.
Vestibular syndromes caused only by cortieallesions
have not yet been weH defined:
Static cortical spatial disorientation may occur as

paroxysmal room-tilt illusion in parietal or


frontal lobe lesions,
contralateral spatial hemineglect in inferior
parietal or frontal lobe lesions,
vertieal neglect below the horizontal meridian in
bilateral parieto-occipitallesions and
tiIts of perceived vertieal (mostly contraversive)
and body lateropulsion in unilateral PIVC lesions
(p. 192).

Dynamic cortieal spatial disorientation with apparent motion or rotational vertigo may occur

in vestibular epilepsy with temporoparietal foci


(p. 233) and
rarely as a transient vertigo in acute lesions of
the vestibular cortex (p. 318).

Spatial hemineglect, a cortical vestibular


syndrome?
Spatial hemineglect impairs focal attention toward
space on the contralesional side. It is most often
induced by acute brain damage of the inferior
parietallobule of the right hemisphere (Vallar and
Perani 1986) and occurs less frequently with acute
right or left lesions of the frontal premotor cortex
(Maeshima et al. 1995; Husain and Kennard 1996). A
recent report described a patient who had sequential
strokes in both hemispheres. After suffering a rightsided parietal infarct, he had a severe unilateral spatial neglect, whieh abruptly disappeared following a
second left-sided frontal infarct (Vuilleumier et al.
1996). Other studies have described single patients
with bilateral inferior parietal lobe lesions, whieh
manifested in vertical neglect of the lower half-space
below the horizontal meridian (Rapcsak et al. 1988;
Shelton et al. 1990). Mesulam {198l} hypothesised

that there is a cortical network for directed attention, in whieh the inferior parietallobule modulates
the shift of attention within extrapersonal space and
the dorsolateral frontal area is responsible for gen erating exploratory motor behaviour. A unilateral
lesion leads to an imbalance of the bilateral tone and
subsequently horizontal displacement of the sagittal
midplane and subjective body orientation toward
the lesioned side.
Studies showing that vestibular (caloric) stimulation significantly improved spatial functioning have
demonstrated the important role of the vestibular
system in neglect (Cappa et al. 1987; Vallar et al.
1993). When vestibular stimulation was combined
with neck muscle vibration, the horizontal deviation
combined linearly, adding or neutralising the effects
observed during application of both types of stimulation (Karnath 1994). This study also showed that
the neglect patients displaced subjective body orientation ipsilesionally, which does not resuIt from a
disturbed primary perception or disturbed transmission of the vestibular or proprioceptive input
from the periphery. Karnath et al. (1991, 1993)
argued that the transformation process converting
the sensory input coordinates from the periphery
into egocentric (body-centred) coordinates is the
critical mechanism that leads to hemineglect: This
process must involve multisensory integration and
motor behaviour including eye and hand movements as weH as walking trajectory (Robertson et al.
1994). Spatial hemineglect also includes the back
space of the body (Vallar et al. 1995). Unilateral
neglect may be limited to visual imagery (Beschin et
al. 1997), but a double dissociation ofvisual imagery
and visual perceptual tasks has also been described
(Coslett 1997). The latter finding suggests that the
disorder is a heterogeneous syndrome attributable
to disruptions of different aspects of spatial cognition.
The concept of critical cortical areas involved in
the control of spatial attention was derived from the
study of patients with circumscribed lesions. Recent
PET studies that visualise the activated neural system underlying visuospatial attention give further
support for this view. Specifically, neocortical activations were observed in the right anterior cingulate
gyrus, the intraparietal sulcus of the right posterior
parietal cortex, and the me sial and lateral premotor
cortices (Nobre et al. 1997).

Paroxysmal room-tilt illusion


Room-tiIt illusions (see also p. 190) are transient
upside-down vision or apparent 90-degree tiIts of
the visual scene. In central vestibular dis orders they

225

Vestibular cortex: its locations, functions, and disorders

can be induced by either acute vestibulocerebellar


brainstem lesions or cortical dysfunction. The two
causative cortical regions are the parieto-occipital
area (Gerstmann 1926; Halpern 1930; Klopp 1951)
and, rarely, the frontal lobe (So1ms et al. 1988). A
lesion in these regions can also cause spatial hemineglect (p. 224). Solms et al. (1988) have carefully
reviewed 21 previously reported cases of transient
upside-down vision and described the following
striking similarities:

''All patients reported the phenomenon of inverted vision during an acute phase in the underlying condition. Indeed, it was often the first sign
of disease.
They usually claimed to experience repeated
episodes in quick succession, but the complaint
was seldom chronic.
The visual phenomenon itself was said to take a
paroxysmal form; that is, the patients reported
that their vision suddenly inverted, remained
upside-down for a few minutes or seconds, and
then rapidly reverted to normal.
Many patients asserted that they were able to
abort the inversion by briefly closing one or both
eyes.
The process of inversion itself was said to take
the form of an actual torsion of the visual scene
in the coronal plane around a central axis.
In all cases in which the direction of this torsion
was noted it was described as clockwise.
All patients reported that their vision rotated
through exactly 180, some described associated
episodes of 90 coronal rotation.
The rotations involved the entire visual array.
Associated proprioceptive changes were frequently no ted, and many patients feIt as if their
bodies were dangling in space.
The subjective experience of inverted vision was
usually accompanied by nausea, vomiting, dizziness, or a general feeling of malaise; nystagmus
was another commonly associated feature.
In all cases in which the site of the lesion was docurnented it involved one of three anatomical
structures: either the parieto-occipital region, the
frontal lobe, or the vestibulocerebellar system:'

Tiliket et al. (1996) reported that a sudden 90 roomtiIt illusion could be elicited in three patients with
unilateral brainstem lesions following vestibular
stimulation by off-vertical rotary chair rotation.
Furthermore, patients with peripheral bilateral
vestibular failure (p. 127) may report transient
room-tiIt illusions on awakening in the morning.
Our latter observation (unpublished) may be com-

parable to the reports of astronauts about occasional


upside-down vision in microgravity (Glasauer and
Mittelstaedt 1992). We believe that room-tiIt illusions
are transient mismatches of the cortical visual and
the vestibular 3-D coordinate maps which occur in
90 or 180 steps as the erroneous resuIt of the
attempted cortical match (Brandt 1998). They should
not be confused with the frequent tiIts of subjective
visual vertical, as they occur with unilateral peripheral vestibular, brainstem, thalamus, or cortex
lesions (p. 189). The matching of two separate sensory 3-D co ordinate maps must be plastic in order to
compensate for visual-vestibular tone imbalances
and/or to adapt to unusual environments (microgravity). The plasticity of this visual-vestibular interaction has been best demonstrated when wearing
reversing prisms (Kohler 1956).

Self-motion perception: the mechanism of


reciprocal inhibitory visual-vestibular
interaction
The vestibular system - a sensor of head accelerations - cannot detect motion at constant velo city
and thus requires supplementary visual information.
Visual perception of self-motion induced by largefield optokinetic stimulation (circularvection, CV;
linearvection, LV; see p. 409) is essential (Dichgans
and Brandt 1978). Vestibular stimuli invariably lead
to the sensation of body motion. Stimuli of visual
motion, however, can always have two perceptual
interpretations: either self-motion or object-motion
(Brandt et al. 1973). The subject who observes moving stimuli may perceive either hirns elf as being
stationary in space (egocentric motion perception)
or the actually moving surroundings as being stable
while he is being moved (exocentric motion perception). Visual self-motion can be perceived while gazing at moving clouds or a train moving on the
adjacent track in a train station. Vestibular information about motion is elicited only through accelerati on or deceleration; it ceases when the cupulae
within the semicircular canals or the otoliths have
returned to their resting position during constant
velocity. Our perception of self-motion during constant velocity car motion is completely dependent on
optokinetically induced vection.
To determine the unknown cortical visualvestibular interaction during CV, we conducted a
PET activation study on CV in human volunteers
(Brandt et al. 1998a). The PET images of activated
cortical areas during optokinetic stimulation without CV were subtracted from those with CV. It was

226

Vertigo

Fig.13.3. PET activation study during large-field optokinetic stimulation inducing circularvection (CV).Comparison ofthe relative
rCBF decreases under both conditions that induce CV (clockwise and counterclockwise) compared to the control condition without CV
(random movement). All voxels shown are significantly above the statistical threshold (p < 0.001 corrected for multiple comparisons) .
The transversal images illustrate deactivation of the posterior insula (parieto-insular vestibular cortex) and ofVS (this deactivation is
only relative to the control condition;compared to baseline, there is an rCBF increase).Thus, the vestibular cortex is deactivated during
visually induced apparent self-motion. An inhibitory visual-vestibular interaction must be assumed during visual self-motion perception. (From Brandt et al. 1998a.)

shown that CV not only bilaterally activates a medial


parieto-occipital visual area separate from motionsensitive areas MT/MST, but simultaneously deactivates the parieto-insular vestibular cortex (Fig. 13.3).
This finding supports a new functional interpretation: reciprocal inhibitory visual-vestibular interaction as a basic sensorimotor mechanism for
adequate self-motion perception. Such a mechanism
protects visual perception of self-motion from
potential vestibular mismatches caused by involuntary head accelerations during locomotion (Brandt
et al. 1998a). This mechanism allows a shift of the
dominant sensorial weight during self-motion perception from one sensory modality (visual or
vestibular) to the other.
Depending on the mode of stimulation, perception of self-motion is dominated by either vestibular
input (head acceleration) or visual input (constant
velo city CV), or both. Quantitative visual-vestibular

interaction is not a simple but rather a complex


process; it refiects not only the pattern of motion
stimulation but also particular active postural and
locomotor tasks. The mechanism of reciprocal
inhibitory visual-vestibular interaction (Fig. 13.4)
makes it possible to relate the dominant perception
of self-motion to the actual input of one of the two
sensory modalities so as to avoid perceptual ambiguity. As a functional consequence, the concurrent
deactivation of the vestibular cortex during CV
should decrease the vestibular system's sensitivity to
head accelerations. This would make the perception
of optokinetically induced CV more robust and
largely insensitive to visual-vestibular mismatches
occurring during involuntary head accelerations in
spatial planes different from the main direction of
locomotion or transportation. The actual horizontal
direction and speed perceived during constant
velo city of car motion are transduced only by the

Vestibular cortex: its locations, functions, and disorders

VIS

PIVC

VIS

PIVC
Fig. 13.4. Schematic representation of reciprocal inhibitory
visual-vestibular interaction as a multisensory mechanism for
self-motion perception. When perception of self-motion is dominated by the vestibular input (accelerations) the visual cortex is
inhibited; this suppresses oscillopsia due to vestibular nystagmus
(top).ln contrast, when perception of self-motion is dominated
by the visual input (e.g., du ring car motion at constant velocity),
the vestibular cortex is inhibited; this suppresses misleading
vestibular inputs caused by involuntary head oscillations during
transportation (bottom). Depending on the mode of stimulation,
this mechanism allows a shift of the dominant sensorial weight
during self-motion perception from one sensory modality (visual
or vestibular) to the other. PIVC = parieto-insular vestibular cortex; VIS = visual cortex; PO = parieto-occipital cortex.

relative optic ftow of the surroundings. Concurrent


vertical vestibular stimulations caused by car motion
and secondary involuntary head accelerations provide vestibular information that is inadequate or
even misleading with respect to self-motion perception in the horizontal direction. It is desirable that
they be suppressed by deactivation of the vestibular
system (Brandt et al. 1998a). This hypothesis is supported by earlier findings that thresholds for detecting vestibular body accelerations (vestibular system)
are significantly increased during optokinetically
induced CV (visual system) in a combined rotatory
chair-drum system (Probst et al. 1985).
In light of these considerations, an earlier observation can be interpreted as the vestibulo-visual
pendant for self-motion perception. In a PET study

227

using caloric vestibular irrigation, activation of the


vestibular cortex caused a significant, bilateral
decrease of rCBF in the occipital visual cortex that
covers Brodmann areas 17,18 and 19 (Wenzel et al.
1996) (Fig. 13.5). Elementary visual hallucinations
have been evoked by caloric stimulation in humans
(Kolev 1995), which might be related to haemodynamic disturbance of the visual system. Using
transcranial Doppler sonography, Tiecks and coworkers (1996) detected areduction of cerebral
blood ftow velo city during caloric vestibular stimulation. This finding suggested to us that deactivation
of the visual cortex is beneficial to the organism during vestibular stimulation, since it suppresses visual
motion input (e.g. distressing oscillopsia, owing to
the retinal slip of the visual scene during vestibular
nystagmus). In the same way that deactivation of the
visual cortex largely protects the vestibular system
from confticting visual motion input, deactivation of
the vestibular cortex prevents optokinetically
induced CV from confticting with vestibular input.
Besides this reciprocal inhibitory interaction, it is
very likely that visual and vestibular cortices have
other forms of interaction depending on the actual
stimulation, the required dynamic spatial orientation, and the intended motor tasks. Furthermore,
somatosensory information about motion must also
be integrated (visual-vestibular-somatosensory
interaction). Activation of both cortices is required
for adequate perception of self-motion and for postural control in stimulus situations involving unexpected, multidirectional transitions between body
accelerations and motion at constant velocity. Both
hemispheres must be activated in stimulus situations in which, for example, both visual fields have
contradictory information about motion. When sitting in a train reading, information about the constant velo city of self-motion is provided by
optokinetic stimulation in one visual hemifield,
whereas the other is filled with contrasts of the stationary train. Since you cannot perceive two different states of body motion at the same time, the two
hemispheres have to correspond to each other and
by an internal shift of attention determine an actual
and unique perception of motion or absence of
motion.
All prestriatal visual areas have transcallosal
interhemispheric connections for this kind of bilateral interaction. In fact, we recently demonstrated
with fMRI that during horizontal optokinetic stimulation the occipito-temporal motion-sensitive areas
MT/MST were activated in both hemispheres of
patients with acute complete homonymous hemianopia (Fig. 13.6). The most likely explanation for
this phenomenon is that the unaffected hemisphere
is activated transcallosally, rather than via ipsilateral

228

Vertigo

Fig.13.5. PET activation/deactivation study following caloric vestibular stimulation. Comparison of the activation conditions versus
rest. Significant decreases of rCBF following vestibular irrigation (n = 6). All significant pixels above the adjusted statistical threshold
(Z > 4.24 in A) and (Z > 4.27 in B) are displayed. All pixels displayed white indicate a Z score> 5. Cold water irrigation of the right ear
(A) and of the left ear (B): top row, from left to right. medial view of the left hemisphere, posterior view, medial view of the right hemisphere; bottom row, lateral view of the corresponding hemispheres and view from above. Shaded areas of the normalised MRI indicate
areas outside the field of view. There is a marked decrease of rCBF restricted to the occipital cortex under both stimulation conditions.
(From Wenzel et al. 1996.)

extrastriatal pathways from the geniculate bodies or


the superior thalamic pulvinar colliculus route
(Brandt et al. 1998). MT/MST provide us with intermediate information ab out visual motion in
extraretinal coordinates, which can be used for

motor control of not only the eyes but also the body.
Interhemispheric interaction requires neural and
interhemispheric synchronisation that is mediated
by cortickocortical connections (Nowak et al. 1995;
Munk et al. 1995).

Fig.13.6. T/-weighted coronal MR images of five cortical sections of a 63-year-old patient with infarction of the right posterior
cerebral artery, which caused complete hemianopia (TRITE = 63/30 ms, CI: = 100). The superimposed activation maps are associated
with optokinetic nystagmus induced by right (a, c, e, g, i) and left motion stimulation (b, d, f. h,j). The colour-coded correlation coefficient scale ranges from 0.5 to a maximum of 1.0. Although motion stimulation was restricted to the left hemisphere, bilateral activation of the motion-sensitive areas MT/MST was found (0; c, h). All other cortical areas including the prefrontal cortex (PF; a,b), the
precentral and posterior median frontal gyrus (frontal eye fields; F; a, b), parts of the parietal cortex including the parietal eye field (P;
a-h), the anterior part (AI; g-j) and posterior part (PI; g-i) of the insula and the primary visual cortex (VC. i-j) were not activated on the
infarcted hemisphere. Thalamic activity was not seen in the infarcted hemisphere, while other subcortical areas such as the putamen,
globus pali idus, caudate nucleus showed bilateral activation. Note that there is no difference in the anatomicallocation and extent of
activation for both directions of object motion. (From Brandt et al. 1998b.)

Vestibular cortex: its locations, functions, and disorders

229

230

References
Akbarian S, Grsser 0-J, Guldin WO (1992) Thalamic connections
of the vestibular cortical fields in the squirrel monkey (Saimiri
sciureus).J Comp NeuroI325:1-19
Akbarian S. Grsser 0-], Guldin WO (1994) Corticofugal connections between the cerebral cortex and brainstem vestibular
nuclei in the macaque monkey. J Comp NeuroI339:421-437
Andersen RA (1987) Inferior parietallobule function in spatial
perception and visuomotor integration. In: Mountcastle VB,
Plum F, Geiger SR (eds) Handbook of physiology. Section I: the
nervous system, vol V. American Physiological Society,
Bethesda, MD, pp 483-518
Andersen RA, Essick GK, Siegel RM (1985) Encoding of spatial
location by posterior parietal neurons. Science 230:456-458
Andersen RA, Gnadt JW (1989) Posterior parietal cortex. In:
Wurtz RH, Goldberg ME, (eds). Reviews in oculomotor
research. 3. The neurobiology of saccadic eye movements.
Elsevier, Amsterdam, pp 315-335
Beschin N, Cocchini G, Della Sala S, Logie RH (1997) Wh at the
eyes perceive, the brain ignores: a case of pure unilateral representational neglect. Cortex 33:3-26
Bisiach E, Rusconi ML, Peretti VA, Vallar G (1994) Challenging
current accounts of unilateral neglect. Neuropsychologia
32:1431-1434
Bottini G, Sterzi R, Paulesu E, Vallar G, Cappa SF, Erminio F,
Passingham RE, Frith CD, Frackowiak RSJ (1994) Identification
of the central vestibular projections in man: a positron emission tomography activation study. Exp Brain Res 99:164-169
Brandt Th (1997) The cortical matching of visual and vestibular
3-D coordinate maps. Ann NeuroI42:983-984
Brandt Th, Dichgans J, Koenig E (1973) Differential effects of central versus peripheral vision on egocentric and exocentric
motion perception. Exp Brain Res 16:476-491
Brandt Th, Btzel K, Yousry T, Dieterich M, Schultze S (1995)
Rotational vertigo in embolic stroke of the vestibular and auditory cortices. Neurology 45:42-44
Brandt Th, Bartenstein P, Janek A, Dieterich M (1998a) Reciprocal
inhibitory visual-vestibular interaction: visual motion stimulation deactivates the parieto-insular vestibular cortex. Brain
121:1749-1758
Brandt Th, Bucher SF, Seelos KC, Dieterich M (1998b) Bilateral
fMRI-activation of motion-sensitive areas MT/MST in
homonymous hemianopia. Arch NeuroI55:1126-1131
Bucher SF, Dieterich M, Seelos KC, Brandt Th (1997) Sensorimotor
cerebral activation during optokinetic nystagmus. A functional
MRI study. Neurology 49: 13 70-13 77
Bucher SF, Dieterich M, Wiesmann M, Weiss A, Zink R, Yousry TA,
Brandt T (1998) Cerebral functional magnetic resonance imaging of vestibular, auditory, and nociceptive areas during galvanic
stimulation. Ann NeuroI44:120-125
Bttner U, Buettner UW (1978) Parietal cortex area 2 V neuronal
activity in the alert monkey during natural vestibular and optokinetic stimulation. Brain Res 153:392-397
Cappa S, Sterzi R, Vallar G, Bisiach E (1987) Remission of hemineglect and anosognosia during vestibular stimulation.
Neuropsychologia 25:775-782
Coslett HB (1997) Neglect in vision and visual imagery: a double
dissociation. Brain 120:1163-117l
Dichgans J, Brandt Th (1978) Visual-vestibular interaction: effects
on self- motion perception and postural contro!. In: R Held,
HW Leibowitz, H-L Teuber (eds) Handbook of sensory physiology, vol VIII Perception, Springer, Berlin Heidelberg New York,
pp 755-804
Dieterich M, Brandt Th (1993) Thalamic infarctions: differential
effects on vestibular function of the roll plan (35 patients).
Neurology 43:1732-1740

Vertigo
Dieterich M, Brandt Th, Bartenstein P, Wenzel R, Danek A, Lutz S,
Ziegler S (1996) Different vestibular cortex areas activated during caloric irrigation: A PET study. J Neurol Suppl 2 243:S40
Dieterich M, Bucher SF, Seelos KC, Brandt Th (1998a) Horizontal
or vertical optokinetic stimulation activates visual motionsensitive, ocular motor and vestibular cortex areas with right
hemispheric dominance. An fMRI study. Brain 121 :1479-1495
Dieterich M, Grnbauer W, Brandt Th (1998b) Direction-specific
impairment of motion perception and spatial orientation in
downbeat and upbeat nystagmus. Neurosci Lett 245:29-32
Faugier-Grimaud S, Ventre J (1989) Anatomic connections of
inferior parietal cortex (area 7) with subcortical structures
related to vestibulo-ocular function in a monkey (Macaca fascicularis). J Comp NeuroI280:1-14
Foerster 0 (1936) Sensible Kortikale Felder. In: Bumke 0, Foerster
o (eds) Handbuch der Neurologie, vol V!. Springer, Berlin
Heidelberg New York, pp 358-449
Fredrickson JM, Figge U, Scheid P, Kornhuber HH (1966)
Vestibular nerve projection to the cerebral cortex of the rhesus
monkey. Exp Brain Res 2:318-327
Friberg L, Olsen TS, Roland PE, Paulson OB, Lassen NA (1985)
Focal increase of blood flow in the cerebral cortex of man during vestibular stimulation. Brain 108:609- 623
Galletti C, Battaglini PP, Fattori P (1993) Parietal neurons encoding spatial orientations in craniotopic coordinates. Exp Brain
Res 96:221-229
Gerstmann J (1926) ber eine eigenartige Orientierungsstrrung
im Raum bei zerebraler Erkrankung. Wien med Wochenschr
76:817-818
Glasauer S, Mittelstaedt H (1992) Determinants of orientation in
microgravity. Acta Astronautica 27:1-9
Grsser O-J, Guldin WO (1995) Primate vestibular cortices and
spatial orientation. In: Mergner T, Hlavacka F (eds) Multisensory control of posture. Plenum Press, New York, pp 51-62
Grsser 0], Pause M, Schreiter U (1990a) Localization and
responses of neurons in the parieto-insular vestibular cortex of
the awake monkeys (Macaca fascicularis). J Physiol
430:537-557
Grsser 0], Pause M, Schreiter U (1990b) Vestibular neurons in
the parieto-insular cortex of monkeys (Macaca fascicularis):
visual and neck receptor responses. J PhysioI430:559-583
Guldin WO, Akbarian S, Grsser 0- J (1992) Cortico-cortical connections and cytoarchitectonics of the primate vestibular cortex: a study in squirrel monkeys (Saimiri sciureus). J Comp
NeuroI324:1-27
Guldin W, Grsser 0- J (1996) The anatomy of the vestibular cortices of primates. In: M. Collard ,M. Jeannerod, Y. Christen (eds)
Le cortex vestibulaire. Ipsen, Boulogne, pp 17-26
Halpern F (1930) Kasuistischer Beitrag zur Frage des
Verkehrtsehens. Z gesamt Neurol Psychiat 126:246-252
Husain M, Kennard C (1996) Visual neglect associated with
frontal lobe infarction. J NeuroI243:652-657
Jeannerod M (1996) Vestibular cortex. A network from directional
coding of behavior. In: M. Collard , M. Jeannerod, Y. Christen
(eds) Le cortex vestibulaire. Ipsen, Boulogne, pp 5-15
Jijiwa H, Kawaguchi T, Watanabe S, Miyata H (1991) Cortical projections of otolith organs in the cat. Acta Otolaryngol (Stockh)
SuppI481:69-72
Jones EG, Burton H (1976) Areal differences in the laminar distribution of thalamic afferents in cortical fields of insular, parietal
and temporal opercular regions of primates. J Comp Neurol
168:197-247
Karnath H-O (1994) Subjective body orientation in neglect and
the interactive contribution of neck muscle proprioception and
vestibular stimulation. Brain 117:1001-1012
Karnath H -0, Schenkel P, Fischer B (1991) Trunk orientation as
the determining factor of the "contralateral" deficit in the
neglect syndrome and as the physical anchor of the internal
representation of body orientation in space. Brain
114:1997-2014

Vestibular cortex: its locations, functions, and disorders


Karnath H-O, Christ K, Hartje W (1993) Decrease of contralateral
neglect by neck muscle vibration and spatial orientation of
trunk midline. Brain 116:383-396
Kawano K, Sasaki M, Yamashita M (1980) Vestibular input to visual
tracking neurons in the posterior parietal association cortex of
the monkey. Neurosci Lett 17:55-60
Klopp H (1951) ber Umgekehrt- und Verkehrtsehen. Deutsch
Zeitschr Nervenheilk 165:231-260
Kohler I (1956) Die Methode des Brillenversuches in der
Wahrnehmungspsychologie mit Bemerkungen zur Lehre der
Adaptation. Z Exp Angew PsychoI3:381-417
Kolev OI (1995) Visual hallucinations evoked by caloric vestibular
stimulation in normal humans. J Vestib Res 5:19-23
Maeshima S, Terada T, Nakai K, Nishibayashi H, Ozaki F, Itakura T,
Komai N (1995) Unilateral spatial neglect due to a haemorrhagic contusion in the right frontal lobe. J NeuroI242:613-617
Mesulam M-M (1981) A cortical network for directed attention
and unilateral neglect. Ann Neurol10:309-325
Mickle WA, Ades HW (1952) A composite sensory projection area
in the cerebral cortex ofthe cat.Am J PhysioI170:682-689
Munk MHJ, Nowak LG, Nelson JI, Bullier J (1995) Structural basis
of cortical synchronization. H. Effects of cortical lesions. J
NeurophysioI74:2401-2414
Nobre AC, Sebestyen GN, Gitelman DR, Mesulam MM,
Frackowiak RSJ, Frith CD (1997) Functionallocalization of the
system for visuospatial attention using positron emission
tomography. Brain 120:515-533
Nowak LG, Munk MHJ, Nelson JI, James AC, Bullier J (1995)
Structural basis of cortical synchronization. I. Three types of
interhemispheric coupling. J Neurophysiol 74:2379-2400
dkvist LM, Schwarz DWF, Fredrickson JM, Hassler R (1974)
Projection of the vestibular nerve to the area 3a arm field in the
squirrel monkey (Saimiri sciureus) Exp Brain Res 21:97-105
Pandya DN, Sanides F (1973) Architectonic parcellation of the
temporal operculum in rhesus monkey and its projection pattern. Z Anat Entwicklungsg 139:127-161
Penfield W, Jasper H (1954) Epilepsy and the functional anatomy
of the human brain. Litde Brown, Boston
Penfield W, Kristiansen K (1951) Epileptic seizure patterns.
Thomas, Springfield, IL
Phillips CG, Powell TPS, Wiesendanger M (1971) Projection from
low threshold muscle afferents of hand and fore arm to area 3a
of baboon's cortex. J Physiol (Lond) 217:419-446
Probst Th, Straube A, Bles W (1985) Differential effects of ambiva-

231
lent visual-vestibular-somatosensory stimulation on the perception of self motion. Behav Brain Res 16:71-79
Rapcsak SZ, Cimino CR, Heilman KM (1988) Altitudinal neglect.
Neurology 38:277-281
Robertson IH, Tegner R, Goodrich SJ, Wilson C (1994) Walking
trajectory and hand movements in unilateral left neglect: a
vestibular hypothesis. Neuropsychologia 32:1495-1502
Schwarz DWF, Fredrickson JM (1971) Rhesus monkey vestibular
cortex: abimodal primary projection field. Science 172:280-281
Schwarz DWF, Deecke L, Fredrickson JM (1973) Cortical projection of group I muscle afferents to areas 2, 3a and the vestibular
field in the rhesus monkey. Exp Brain Res 17:516-526
Shelton PA, Bowers D, Heilman KM (1990) Peripersonal and
vertical neglect. Brain 113:191-205
Smith BH (1960) Vestibular disturbance in epilepsy. Neurology
10:465-469
Solms M, Kaplan-Solms M, Saling M, Miller P (1988) Inverted
vision after frontal lobe disease. Cortex 24:499-509
Tiecks FP, Planck J, Haberl RL, Brandt T (1996) Reduction in posterior cerebral artery blood flow velo city during caloric
vestibular stimulation. J Cerebr Blood Flow Metab
16:1379-1382
Tiliket C, Ventre-Dominey J, Vighetto A, Grochowicki M (1996)
Room tilt illusion. A central otolith dysfunction. Arch Neurol
53:1259-1264
Vallar G, Perani D (1986) The anatomy of unilateral neglect after
right hemisphere stroke lesions: a clinical CT correlation study
in man. Neuropsychologia 24:609-622
Vallar G, Bottini G, Rusconi ML, Sterzi R (1993) Exploring
somatosensory hemineglect by vestibular stimulation. Brain
116:71-86
Vallar G, Guariglia C, Nico D, Bisiach E (1995) Spatial hemineglect
in back space. Brain 118:467-472
Vuilleumier P, Hester D, Assal G, Regli F (1996) Unilateral spatial
neglect recovery after sequential strokes. Neurology 19: 184-189
Walzl EM, Mountcasde VB (1949) Projection of vestibular nerve to
cerebral cortex of cat. Am J Physiol 159:595
Watt DGD (1997) Pointing at memorised targets during prolonged microgravity. Aviat Space Environm Med 68:99-103
Wenzel R, Bartenstein P, Dieterich M, Danek A, Weindl A,
Minoshima S, Ziegler S, Schwaiger M, Brandt Th (1996)
Deactivation of human visual cortex during involuntary ocular
oscillations: A PET activation study. Brain 119:101-110

Vestibular epilepsy

Vestibular epilepsy (vestibular seizures or auras) is a


rare cortical vertigo syndrome secondary to focal
epileptic discharges in either the temporal lobe or
the parietal association cortex (Foerster 1936;
Penfield and Jasper 1954; Schneider et al. 1968).
Multiple areas of both receive bilateral vestibular
projections from the ipsilateral thalamus. Results of
electrical stimulation in humans have shown that the
intraparietal sulcus (Foerster 1936) and the posterior part of the superior temporal gyrus (Penfield and
lasper 1954; Penfield 1957) form part of the vestibular cortex, as described in Chap. 13 (po 219). Similarly
experiments in monkeys have identified vestibular
neuronal activity in the posterior caudal part of the
postcentral gyrus (Fredrickson et a1. 1966) and in
the deep posterior insula (parieto-insular vestibular
cortex, PIVC; Grsser et a1. 1990a,b). It is not yet
known which of the different multisensory vestibular cortical areas (2v, 3a V, 6, 7, PIVC) is involved in
vestibular epilepsy. If vestibular seizures arise from
different areas, the sensorimotor symptomatology
may differ as regards apparent rotation or tilt (Smith
1960), with or without associated eye, head and body
deviation or epileptic nystagmus. Clinical data on
the directions of apparent self-motion or surroundmotion are mostly incomplete and imprecise. If the
description is exact, as in rota tory seizures ("volvular epilepsy"), then the topographic localisation of
the underlying pathology is too inexact to per mit its
allocation to known vestibular areas.
An acute unilateral functional deficit of the
vestibular cortex (for example, in medial cerebral
artery infarction) rarely manifests with vertigo
(Brandt et al. 1995), unlike lesions in the vestibular
area of the brainstem. It is not the functionalloss but
the focal discharge that causes central vertigo. This
has been repeatedly demonstrated by stimulation
experiments (Fig. 14.1). Electrical stimulation of the
human thalamus during stereotactic neurosurgical
procedures induced sensations of movement in
space. They were most frequently described as

horizontal or vertical rotation or as sensations of


falling or rising (Hawrylyshyn et al. 1978). These
sensations were similar to tho e induced by stimulation of the vestibular cortex (Foerster 1936; Penfield
and lasper 1954).
Vertigo has long been considered a manifestation
of epileptic auras (Jackson 1879; Gowers 1907). Most
information on auras (ve tibular epilepsy), including
case descriptions, comes from older textbooks, for
example Bumke and Foerster (1936) or Penfield and
Jasper (l954), or review articles (Penfield and
Kristiansen 1951; Karbowski 1984). Original studies
are relatively rare (Riser et a1. 1951; Pedersen and
Jepson 1956; Hughes and Drachman 1977;
Kogeorgos et al. 1981) in contrast to studies of the
less important phenomenon of vestibulogenic
epilepsy (p. 238).
Knowledge of vestibular epilepsy has not
advanced very much in the past 40 years because of
the lack of pathophysiological data. The limited clinical data that exist consist of patients' descriptions,
EEG recordings and correlations of vestibular symptoms with the results of brain imaging techniques or
surgical findings. Furthermore, the syndrome of
vestibular epilepsy is not always weil defined. Vertigo
and dizziness in the sense of"lightheadedness" are
often described by patients as part of the aura in
complex partial seizures with temporal or extratemporal foci (Fried et al. 1995). Most patients with
(simple partial) vestibular seizures also have complex partial seizures or generalised tonic-clonic
seizures (Kogeorgos et a1. 1981). Vestibular seizures
must be distinguished from versive seizures that
manifest with contralateral rotation of head and eye.
The latter are typically associated wirh either motor
cortex foci or are the result of seizure spread from
the temporal lobes.
Epileptic nystagmus usually beats contraversive to
the seizure focus and may be of vestibular, visual
(optokinetic), or cortical ocular motor origin
(Kaplan and Tusa 1993).

233

234

The vestibular seizure


Vestibular seizures are auras (simple or complex
partial sensory seizures) with vertigo as the predominant symptom. The patient experiences sudden
disequilibrium with rotational (horizontal or vertical) or linear (rising, falling, tilting) vertigo, accompanied in most cases by body or head and eye
rotation (with or without nystagmus). The symptoms may last a few seconds or some minutes, and
may be associated with mild nausea (vomiting is not
typicai). Tinnitus, often unilateral and contralateral
paraesthesia (but less frequently olfactory or gustatory sensations) may precede or accompany the vertigo, which is indicative of a lateral and superior
rather than medial and basal temporallocation of
the focus (Fig.14.1).
From the few detailed reports on the direction of
apparent self-motion and surround motion, it is
most likely that the direction of perceived selfmotion, measurable body motion, and eye deviation
is contraversive to the epileptic focus, whereas
simultaneously perceived surround motion may be
ipsiversive (Foerster 1936). The different effects on
the perception of self-motion and object-motion
correspond to the perception of normallocomotion,
in which body motion in one direction causes relative surround motion in the opposite direction. The
finding that the directions of nystagmus and of measurable body fall are the same differs significantly
from the situation in peripheral vestibular vertigo
syndromes, in which the directions of nystagmus
and body fall are always opposite (p. 14). Thus, in
vestibular epilepsy actual body movements do not
represent a vestibulospinal compensation of perceived vertigo but an epileptic postural response.
Vestibular seizures can manifest without any objective eye or body movements, as described by
Foerster (1936) in his stimulation experiments. In
the case described by Penfield and Jasper (1954), the
direction of eye deviation, nystagmus and perceived
surround motion was the same; however, this case
appears to represent visual occipital lobe (p. 237)
rather than vestibular epilepsy.

Rotatory seizure ("volvular epilepsy")


The rare rotatory seizures ("volvular epilepsy", cirding epilepsy) are characterised by paroxysmal,
repetitive walking in small cirdes without impairment of consciousness (Schneider et al. 1968, 1971;
Remillard et al. 1974; Hochman 1983; Donaldson
1986; Schiffter 1987; Pohlmann-Eden and
Daffertshofer 1994). When the direction of rotation

Vertigo

was reported, it was usually contraversive to the


epileptic focus; however, in two of the 13 described
cases, the direction was ipsiversive (Schneider et al.
1971; Pohlmann-Eden and Daffertshofer 1994). The
circling, which can continue for several complete
turns, is sometimes preceded by versive movements
of head and body in the same direction. This may
develop into secondarily generalised tonic-donie
seizures (Pohlmann-Eden and Daffertshofer 1994).
The localisation of the focus and the pathology was
either frontal or temporal. Hs association with various vestibular, auditory, or autonomous auras makes
it probable that the temporal lobe is involved. Some
of the patients reported vertigo and dizziness;
nystagmus has not been observed.

Differential diagnosis
Differential diagnosis of vestibular seizures indudes
other epileptic seizures such as limbic complex partial seizures, versive seizures, visual seizures, or
absences, the non-epileptic vestibular paroxysmia
(p. 117), or paroxysmal dysarthria/ataxia, which is a
manifestation of multiple sclerosis (p.242).
Vestibular seizures can be easily distinguished from
vestibular dis orders such as Meniere's disease or
vestibular neuritis by their short duration and the
other types of epileptic seizures seen in these
patients. They are not precipitated by changes in
head position, which excludes all positional vertigo
syndromes (p. 247). Vertigo in migraine, for example, benign paroxysmal vertigo of childhood,
basilar migraine (p. 329), transient vertebro-basilar
ischaemia (p. 307), familial episodie ataxia (p. 365),
or phobie postural vertigo (p. 469) - all can mimie
vestibular epilepsy. Vestibular seizures must be further distinguished from atonic seizures (Kolbinger
and Jerusalem 1994), cataplectic attacks in narcolepsy, hyperexplexia ("startle disease") and drop
attacks of vestibular or non-vestibular origin
(p. 15). Drop attacks (Meissner et al. 1986) can occur
with both endolymphatic hydrops (p. 94) and
epilepsy, but they must not be confused with
vestibular epilepsy.
While an abnormal interictal EEG with focal slowing or sharp waves over temporoparietal regions
may be very helpful for diagnosis, anormal EEG
does not exdude vestibular epilepsy. Whenever possible, an ictal polygraphie video-EEG recording
should be obtained for differential diagnosis.
Epilepsy and a vestibular vertigo syndrome can
coincide by chance, or both can be due to the same
causative factor such as head trauma, meningoencephalitis, or vascular disease (Karbowski 1984).
Vertigo as an adverse effect of antiepileptic therapy

Vestibular epilepsy

235

~ = motor left hand


() = motor left arm
E!) = sensory left hand
0 = ocular motor
= motor tongue
~
@

= motor throat
= somatosens. aura

= rotatory vertigo to the left,


acoustic hallucinations

= head deviation to the left

= motor head

Fig.14.1. Sensory and motor effects of electrical stimulation by subdural platinum electrodes in a 22-year-old patient with right
perirolandic epilepsy due to cortical dysplasia. Stimulation of electrode 11 resulted in contraversive rotatory vertigo at low stimulation
intensities.ln addition, acoustic hallucinations were consistently elicited with higher stimulation intensities at the same electrode, indicating concurrent excitation of Heschl's transverse gyrus and the homologue of the parieto-insular vestibular cortex in the posterior
insula. (Courtesy of Dr S. Noachtar, Department of Neurology, University of Munich, Germany.)

is not a paroxysmal attack but rather a continuous


sensation (p. 395).
The aetiology of vestibular epilepsy is the same as
that of all other focal epilepsies, such as congenital
lesions, trauma, haemorrhage, ischaemia, encephalitis,
and most important, tumours.
Table 14.1 summarises the information given in
this chapter about vestibular epilepsy.

Management
Vestibular seizures respond to antiepileptics. Firstline drugs are carbamazepine or phenytoin; secondline drugs are gabapentine, sodium valproate and
lamotrigine (Schmidt and Shorvon 1996). If necessary and possible, surgical procedures may be considered (Noachtar et al. 1996).

Epileptic nystagmus
After several cases of epileptic nystagmus were
reported to be apart of minor or major seizures in

the French, Italian and German literature (for example, temporal lobe epilepsy: Arend et al. 1968;
grand mal: Stauder 1934), von Rad (1970), White
(1971) and Furman et al. (1990) described isolated
cases of epileptic nystagmus not associated with
other more typical seizure manifestations. This ictal
nystagmus is induced by occipital or temporoparieto-occipital epileptogenic zones; it is either a
contraversive, rotatory, horizontal jerk (von Rad
1970; Furman et al. 1990) or it changes in its beating
direction (Stolz et al. 1991) or beating characteristics from jerk to pendular (White 1971). Cortical
blindness is a well-recognised post-ictal phenomenon (Sadeh et al. 1983), but it has also been
described as an ictal manifestation in occipitallobe
epilepsy (von Rad 1970; Huott et al. 1982). Seizures
arising from the occipitallobe which are induced by
the neurotoxicity of cisplatin were characterised by
cortical blindness, ictal (direction changing) horizontal jerk nystagmus, eyelid twitching (without
optokinetic response) and a confusional state
(Kattah et al. 1986). Not only clouding of vision but
also apparent jumping or movement of the visual
scene may be experienced.
Thus, epileptic nystagmus is not at all a specific
indicator of epileptogenic zones in the vestibular

Vertigo

236
Table 14.1. Vestibularepilepsy
Clinical syndrome

- Sudden rotational or linear vertigo aeeompanied by eye-head-body


deviation or rotation (contraversive to lesion) lasting for some seconds
or minutes; tinnitus or contralateral paraesthesia may be assoeiated;
nausea is moderate or absent
(rare rotatory seizures with walking in "small circles")
- Contraversive nystagmus may be present but ean also be due to "visual
(optokinetie) epilepsy"
- Abnormal EEG patterns and lateral temporal epileptie foei are frequent
Incidencel/age/sex

- Depends on aetiology, no preponderanee in either sex

Pathomechanism

- Epileptie discharges of vestibular cortex


- Struetural or funetionallesions predominantly (?) involve the parietoinsular vestibular cortex
Aetiology

- Congenitallesions, tumours, haemorrhage, vaseular isehaemia, posttraumatie, eneephalitis, intoxieation (cis platin)

Table 14.2. Overview of the eharaeteristies of epileptie nystagmus


(based on a total of 51 eases described in the literature up to 1996)

Direetion of nystagmus (n = 51)


Contraversive
Ipsiversive
Upbeat
"Vertieal"
Pendular
Changing direetion
Poorly deseribed
Associated version of eye and/or head (n = 27)
Toward fast phase
Toward slow phase
Loealisation of epileptie foeus and/or pathology (n = 51)
Tem poro- pa rieto-oeei pita I
Frontal
Temporal
Generalised/diffuse

37
1
2
4

21
6

39
6
2

Course/prognosis

- Depends on aetiology: in most patients simple partial vestibular seizures


evolve into complex partial and/or secondarily generalised seizures
Management

- Effeetive medieal treatment with anticonvulsants (earbamazepine,


phenytoin)
- Surgieal therapy of foeal epilepsy
Differential diagnosis

- Epileptic: limbie complex partial, visual, versive, or atonie seizures,


absenees
- Non-epileptic: recurrent vertigo attaeks in migraine, vertebrobasilar
isehaemia, drop attaeks (vestibular and non-vestibular), paroxysmal
attaeks in multiple sclerosis, psyehogenie disorders, episodie (familial)
ataxia
"Vestibulogenic epilepsy"

- This eondition should not be confused with vestibular epilepsy; it is a


variety of sensory-evoked epilepsy indueed by peripherallabyrinthine
stimulation (ealorie irrigation or spinning)
Epileptie nystagmus (non-vestibular)
Mostly contraversive horizontal nystagmus in the awake patient due to
- epileptie aetivation of eortieal saeeade generating centres
- epileptie aetivation of cortieal smooth pursuit generating centres
- visual (optokinetie) seizures with apparent motion of the visual scene

cortex; it can ongmate in the vestibular, visual


(optokinetic), or oculomotor cortex. It may represent only cortical ocular motor activation. Perfusion
studies of an infant with epileptic nystagmus
demonstrated interictal hypoperfusion and ictal
hyperperfusion of the left temporal lobe, showing
the cortical origin. This also indicates that the infant
cortex has functioning efferent connections to the
brainstem oculomotor centres at 10 days of age
(Harris et al. 1997). Epileptic nystagmus was reported to occur with ipsiversive eye deviation and was
consequently attributed to the cerebral smooth pur-

suit pathways (Tusa et al. 1990). However, contraversive ocular deviations and nystagmus, which
involved cortical control of saccadic eye movements
(Kaplan and Lesser 1989; Kaplan and Tusa 1993), are
much more frequent. Kaplan and Tusa (1993)
describe eight patients with horizontal epileptic
nystagmus and hypothesise that the frequency of
ictal discharge, anatomic localisation of ictal activity
and level of consciousness determine its occurrence
and mechanism. Both quick and slow phases were
confined to the "Schlagfeld" contraversive to the
seizure focus.
Kaplan and Tusa (1993) discuss three possible
mechanisms for epileptic nystagmus in awake
patients:
epileptic activation of cortical saccade regions,
causing contraversive quick phases combined
with a defect in the gaze-holding system allowing
the eyes to drift backward toward the midline;
2. epileptic activation of cortical pursuit regions,
generating ipsiversive slow phases combined
with quick phases that are reflexively generated;
3. epileptic activation of cortical optokinetic
regions.
1.

f 51 patients described in the literature, epileptic


nystagmus was contraversive in 37, ipsiversive in one
(Table 14.2). The pathology was in most cases roughly
localised in the temporoparieto-occipital area (Stolz
et al. 1991).
Epileptic nystagmus or ocular oscillations may
also occur during absence seizures (Watanabe et al.
1984; Yaqub 1993), and may be associated with
pupillary oscillations in complex partial seizures
(Lavin 1986). Simon and Aminoff (1986) have

237

Vestibular epilepsy

described small periodic vertical eye movements in


anoxic coma as the sole clinical manifestation of
status epilepticus. Kaplan and Lesser (1989) reported
upbeat nystagmus with upward deviation of the eyes
associated with low voltage (3-4 Hz) bifrontal spikes
in a comatose patient. It is debatable whether the
divergent-convergent eye movements and transient
eyelid openings associated with an EEG burstsuppression pattern in a neonate represent a brainstern release phenomenon (Nelson et al. 1986) or
epileptic nystagmus (Jacome 1986). Viguaendra and
Lim (1978) reported on a patient with focal epilepsy
and secondarily generalised seizures in whom convergence elicited seizures in the right occipital area
as a kind of"reflex epilepsy". Finally, curious epileptic eye movements such as monocular ictal nystagmus (Jacome and Fitzgerald 1982) cannot be
explained simply by cortical activity, since supranuclear eye movements are usually conjugate.

Vestibular versus visual (optokinetic)


seizures
As early as 1936, Foerster discussed the question of
wh ether vertigo in epilepsy is due to direct excitation
of the sensory (vestibular) cortex or a kind of perceptual consequence of excitation of the oculomotor cortex. Jackson (1879) believed that ictal vertigo was a
manifestation of a motor cortex dis charge in the
frontal lobe. From his intrasurgical stimulation experiments, Foerster inferred that vertigo in focal epilepsy
can be due simply to sensory vestibular cortex activation in which severe rotational vertigo is induced
without concurrent ocular deviation or nystagmus.
On the other hand, so-called vestibular seizures can
be associated with nystagmus, and mostly manifest
with some rotation of body or head and eye.
We have observed an 11-year-old girl who experienced horizontal oscillopsia of ambient (peripheral)
vision with stationary metamorphopsia of focal
(parafoveal) vision during visual auras. This correlated with horizontal nystagmus to the left and
right temporo-occipital EEG discharges (Stodieck et
al. 1990). During the seizure there was no impairment of consciousness. The girl could look around
freely during the ictal nystagmus. In darkness she
experienced oscillating visual hallucinations and a
central stationary fortification spectrum. The only
neurological signs in the interval were a slight
smooth pursuit defect and right temporo-occipital
photo sensitive spikes and spike waves. We concluded
that contraversive ictal nystagmus without concomitant vertigo and postural imbalance occurs as a

simple partial visual seizure due to excitation of the


non-striate visual movement detection areas
(Stodieck et al. 1990). This causes hallucination of
egocentric object-motion within the visual scene as
opposed to the experience of self-motion in vestibular epilepsy. The hallucinatory motion is pursued as
if it is real (see hypnotic nystagmus; Brady and Levitt
1964), and thus represents a kind of"epileptic optokinetic nystagmus". The posterior cortical areas may
involve the middle temporal (MT) or medial superior
temporal (MST) areas, which are important for the
visual processing required for ocular tracking of a
moving target and which - if lesioned - impair
ipsiversive smooth pursuit (Thurston et al. 1988) and
OKN (Kmpf 1986).
"Visual (optokinetic) seizure" of occipitotemporal
origin must be differentiated from vestibular
seizures of temporoparietal origin. In typical
vestibular seizures the patient experiences apparent
self-motion and surround motion with ipsiversive
head and eye rotation, with or without nystagmus.
Video-EEG and analysis of perceptual, oculomotor,
postural, and vegetative symptoms per mit a clinical
differentiation.

"Vestibulogenic epilepsy"
Vestibular epilepsy should be distinguished from
"vestibulogenic epilepsy". Early clinical observation,
carried out before the invention of EEG and before
modern experimental studies, provided evidence
that vestibular dysfunction or stimulation could
induce seizures (Gowers 1907; Marie and Pierre
1922; Stauder 1934; Spiegel 1932, 1934). The term
"vestibulogenic epilepsy", coined by Behrman and
Wyke (1958), was used by others (Orban and Lang
1963) to describe a variety of sensory-evoked epilepsies, in which seizures were induced by peripheral
labyrinthine stimulation (caloric irrigation or body
rotation). The term "reflex epilepsy" was also used.
But, as Barac (1968) points out, this condition is
rather non-specific for two reasons: (1) it is only
occasionally possible to provoke the seizures by
labyrinthine stimulation (seizures also occur spontaneously), and (2) vertigo is not at all a typical
manifestation of the seizures, which may be complex
partial or secondarily generalised tonic-clonic. The
term "vestibulogenic epilepsy", therefore, is somewhat vague and should be used with caution.
Molnar et al. (1959), Barac (1968) and Cantor
(1971) showed that caloric irrigation can suppress or
activate paroxysmal EEG-patterns, or activate a
focus that was not visible before. While vestibular

238

stimulation (caloric irrigation, body rotation) was


repeatedly used as a diagnostic EEG-provocation
test in epilepsy, it certainly cannot be regarded as a
useful method. The asymmetrical changes in alpha
rhythm in response to vestibular stimulation
(Karbowski 1971; Barac 1967) critically depend on
the direction of induced nystagmus and therefore
may represent ocular motor activity, which is unrelated to epileptic discharges. Furthermore, arecent
PET study (Wenzel et al. 1996) demonstrated that
caloric irrigation deactivates visual cortex activity
(p.225), and a transcranial Doppler sonography
study showed that it reduces posterior cerebral
artery blood ftow velo city (Tiecks et al. 1996).

References
Arend R, Brzacki A, Misztal S (1968) Nystagmusanflle bei
Temporallappenepilepsie. Psychiat Neurol Med Psychol (Lpz)
20:273-276
Barac B (1967) Vestibular influence upon the EEG of epileptics.
Electroenceph Clin NeurophysioI22:245-252
Barac B (1968) Vertiginous epileptic attacks in so-called "vestibulogenic seizures" . Epilepsia 9: 13 7-144
Behrman S, Wyke BD (1958) Vestibulogenic seizures. A consideration of vertiginous seizures with particular reference to convulsion produced by stimulations of labyrinthine receptors. Brain
81:529-541
Brady JP, Levitt EEC (1964) Nystagmus as a criterion ofhypnotically induced visual hallucinating. Science 146:85-86
Brandt Th, Btzel K, Yousry T, Dieterich M, Schultze S (1995)
Rotational vertigo in embolic stroke of the vestibular and auditory cortices. Neurology 45:42-44
Bumke 0, Foerster 0 (1936) (eds) Handbuch der Neurologie, vol
VI. Springer, Berlin.
Cantor FK (1971) Vestibular-temporallobe connections demonstrated by induced seizures. Neurology 21:507-516
Donaldson IM (1986) Volvular epilepsy: a distinctive and underreported seizure type. Arch NeuroI43:260-262
Foerster 0 ( 1936) Sensible kortikale Felder. In: Bumke 0, Foerster
o (eds) Handbuch der Neurologie, vol VI. Springer, Berlin, pp
358-448
Fredrickson JM, Figge U, Scheid P, Kornhuber HH (1966)
Vestibular nerve projection to the cerebral cortex of the rhesus
monkey. Exp Brain Res 2:318-327
Fried I, Spencer DD, Spencer SS (1995) The anatomy of epileptic
auras: focal pathology and surgical outcome. J Neurosurg
83:60-66
Furman JMR, Crumrine PK, Reinmuth OM (1990) Epileptic nystagmus.Ann NeuroI27:686-688
Gowers WR (1907) The borderlands of epilepsy. Churchill,
London
Grsser OJ, Pause M, Schreiter U (1990a) Localization and
responses of neurons in the parieto-insular vestibular cortex of
the awake monkeys (Macaca fascicularis). J Physiol
430:537-557
Grsser OJ, Pause M, Schreiter U (1990b) Vestibular neurons in
the parieto-insular cortex of monkeys (Macaca fascicularis):
visual and neck receptor responses. J PhysioI430:559-583
Harris CM, Boyd S, Chong K, Harkness W, Neville BGR (1997)
Epileptic nystagmus in infancy. J Neurol Sci 151:111-114

Vertigo
Hawrylyshyn PA, Rubin AM, Tasker RR, Organ LW, Fredrickson
JM (1978) Vestibulothalamic projections in man - a sixth
primary sensory pathway. J NeurophysioI41:394-401
Hochman MS (1983) Rotary seizures associated with frontal lobe
malignant neoplasm: a case report. Epilepsia 24: 11-14
Hughes MR, Drachman DA (1977) Dizziness, epilepsy and EEG. J
Nerv Ment Dis 38:431-435
Huott AD, Madison DS, Niedermeyer E (1982) Occipital lobe
epilepsy. Europ Neurolll:325-339
Jackson H (1879) Diagnosis of epilepsy. Medical Times and
Gazette 1:29
Jacome DE (1986) Epileptic nystagmus and eye movements. J Clin
Neuro-OphthaI6:269-270
Jacome DE, Fitzgerald R (1982) Monocular ictal nystagmus. Arch
Neurol 39:653-656
Kaplan PW, Lesser RP (1989) Vertical and horizontal epileptic
gaze deviation and nystagmus. Neurology 39: 1391-1393
Kaplan PW, Tusa RJ (1993) Neurophysiologie and clinical correlations of epileptic nystagmus. Neurology 43:2508-2514
Karbowski K (1971) Vestibularapparat und hirnelektrische
Aktivitt. Hans Huber, Bern
Karbowski K (1984) Schwindel und Epilepsie. Therapeut
Umschau 41:705-708
Kattah JC, Potolicchio J, Kotz HL, Kolsky MP, Thomas D (1986)
Cortical blindness and occipitallobe seizures induced by cisplatin. Neuro-ophthalmology 7:99-104
Kogeorgos J, Scott DF, Swash M (1981) Epileptic dizziness. Br Med
J 282:687-689
Kolbinger HM, Jerusalem F (1994) Sturzanflle. Akt Neurol 21:2-8
Kmpf D (1986) The significance of optokinetic nystagmus asymmetry in hemispheric lesions. Neuro-ophthalmology 6:61-64
Lavin PJM (1986) Pupillary oscillations synchronous with ictal
nystagmus. Neuro-Ophthalmol 6: 113-116
Marie P, Pierre JR (1922) Etudes ru la variabilite des reactions
vestibulaires des epileptiques etudiees par la methode de
Baniny. Rev NeuroI38:86-92
Meissner J, Wiebers DO, Swendson JW, O'Fallon WM (1986) The
natural history of drop attacks. Neurology 36:1029-1034
Molnar L, Kekesi F, Gosztonyi G (1959) Die Wirkung der
Labyrinthreizung auf das normale und pathologische
Elektroencephalogramm beim Menschen. Arch Psychiat Z Ges
NeuroI199:152-171
Nelson KR, Brenner RP, Carlow TI (1986) Divergent-convergent
eye movements and transient eyelid opening associated with an
EEG burst-suppression pattern. J Clin Neuro-ophthaI6:43-46
Noachtar S, Lders HO, Bromfield EB (1996) Surgical therapy of
epilepsy. In: Brandt Th, Caplan LR, Dichgans J, Diener HC,
Kennard C (eds) Neurological dis orders: course and treatment.
Academic Press, San Diego, pp 183-191
Orban L, Lang J (1963) Zur Pathogenese der vestibulogenen
Epilepsie. Psychiat Neurol (Basel) 146:193-198
Pedersen E, Jepson 0 (1956) Epileptic vertigo. Acta Psychiatr
Neurol Scand (Suppl) 108:301-310
Penfield W (1957) Vestibular sensation and the cerebral cortex.
Ann Oto-Rhino-Laryngol (St Louis) 66:691-698
Penfield W, Jasper H (1954) Epilepsy and the functional anatomy
of the human brain. Liltle Brown, Boston
Penfield WG, Kristiansen K (1951) Epileptic seizure patterns.
Thomas, Springfield, IL.
Pohlmann-Eden B, Daffertshofer M (1994) Rotatorische Anflle:
ein seltenes quivalent fokaler epileptischer Aktivitt.
Nervenarzt 65:415-420
Rad von M (1970) Ein Fall von isoliertem epileptischem
Nystagmus. Dtsch Z Nervenheilk 197: 125-132
Remillard GM, Ethier R, Andermann F (1974) Temporal lobe
epilepsy and perinatal occlusion of the posterior cerebral
artery. Neurology 24: 1001-1009
Riser M, Geraud J, Grezes-Rueff Ch, Lavitry S, Rascol A (1951) A
propos de 14 cas d'epilepsie giratoire. Rev NeuroI85:245-253

Vestibular epilepsy
Sadeh M, Goldhammer Y, Kuritsky A (1983) Postictal blindness in
adults. J Neurol Neurosurg Psychiatry 46:566-569
Schiffter R (1987) Kreislaufen als Ausdruck fokaler epileptischer
Aktivitt. Akt NeuroI14:132-133
Schmidt D, Shorvon S (1996) The epilepsies. In: Brandt Th, Caplan
LR, Dichgans J, Diener HC, Kennard C (eds) Neurological disorders: course and treatment. Academic Press, San Diego, pp
159-181
Schneider RC, Calhoun HD, Crosby EC (1968) Vertigo and rotational movement in cortical and subcorticallesions. J Neurol
Sei 6:493-516
Schneider RC, Calhoun HD, Kooi KA (1971) Circling and rotational
automatisms in patients with frontotemporal cortical and subcorticallesions. J Neurosurg 35:554-563
Simon RP, Aminoff MJ (1986) A clinical sign predicting electroencephalographic status epilepticus in anoxic coma. Neurology
36, (Suppl) 1:287
Smith BH (1960) Vestibular disturbance in epilepsy. Neurology
10:465-469
Smith NJ, Docherty TB (1982) Nystagmus: an unusual manifestation of temporal lobe epilepsy. J Electrophysiol Tech 8:7-13
Spiegel EA (1932) Rindenerregung (Auslsung epileptiformer
Anflle) durch Labyrinthreizung. Versuch einer Lokalisation
der corticalen Labyrinthzentren. Z Ges Neurol Psychiat
138: 178-196
Spiegel EA (1934) Labyrinth and cortex. The electroencephalogram of the cortex in stimulation of the labyrinth. Arch Neurol
Psychiatr (Chic) 31:469-482

239
Stauder KH (1934) Epilepsie und Vestibularapparat. Arch Psychiat
Nervenkr 101:739-761
Stodieck SRG, Brandt Th, Bttner U (1990) Visual and vestibular
epileptic seizures. Electroenceph Clin Neurophysiol 75:65-66P
Stolz SE, Chatrian GE, Spence AM (1991) Epileptic nystagmus.
Epilepsia 32:910-918
Thurston SE, Leigh RJ, Crawford T, Thompson A, Kennard C
(1988) Two distinct defieits ofvisual tracking caused by unilateral lesions of cerebral cortex in humans. Ann Neurol
23:276-273
Tiecks FP, Planck J, Haber! RL, Brandt T (1996) Reduction in posterior cerebral artery blood flow veloeity during caloric
vestibular stimulation. J Cerebr Blood Flow Metab
16:1379-1382
Tusa RJ, Kaplan PW, Hain TC, Naidu S (1990) Ipsiversive eye deviation and epileptic nystagmus. Neurology 40:662-665
Viguaendra V, Lim CL (1978) Epileptic discharges triggered by eye
convergence. Neurology 28: 589-591
Watanabe K, Negoro T, Matsumato A, Inokuma K, Takaesu E,
Machara M (1984) Epileptic nystagmus assoeiated with typical
absence seizures. Epilepsia 25:22-24
Wenzel R, Bartenstein P, Dieterich M, Danek A, Weindl A,
Minoshima S, Ziegler S, Schwaiger M, Brandt Th (1996)
Deactivation of human visual cortex during involuntary ocular
oscillations: a PET activation study. Brain 119:101-110
White JC (1971) Epileptic nystagmus. Epilepsia 12:157-164
Yaqub BA (1993) Electroclinical seizures in Lennox-Gastaut syndrome. Epilepsia 34:120-127

Miscellaneous central vestibular


disorders

The typical central vestibular syndromes have been


described in Chapters 10- 14. They are related to the
central vestibular structure affected, but do not have
a specific aetiology. Since a vascular aetiology seems
most important clinically, stroke and vertigo (Chap.
19) and migraine and vertigo (Chap. 20) have been
described separately. Other central vestibular conditions that are also described in separate chapters are
central positional vertigo (Chap. 18), familial episodic
ataxia (Chap. 25), drugs and vertigo (Chap. 28), and
vestibular dysfunction in childhood (Chap. 26) or in
old age (Chap. 27). This chapter deals with certain
aspects that have not been sufficiently covered elsewhere. These include central brainstem lesions mimicking peripheral vestibular disorders, forms of
paroxysmal central vertigo, falls without vertigo,
vestibular syndromes associated with multiple
sclerosis or brain tumours, and the question of
metabolic dis orders and vestibular dysfunction.

Central brainstem/cerebellar lesions


mimicking vestibular neuritis or
peripheral vestibular failure
Small cerebellar infarctions in the territory of either
the anterior inferior cerebellar artery (AI CA) or the
posterior inferior cerebellar artery (PI CA) have been
repeatedly reported to mimic unilateral peripheral
labyrinthine dis order (Chap. 19), and to generally
take a similarly benign course (Duncan et al. 1975;
Guiang and Ellington 1977; Rubinstein et al. 1980;
Samson et al. 1981; Kmpf 1986; Disher et al. 1991).
Since both the areas covered by these arte ries and
the anatomy of the collaterals vary it is sometimes
difficult to decide which artery is responsible for
clinical signs or infarcted areas seen on MRI. When
the PI CA is hypoplastic, the AICA takes over the territory usually supplied by the lateral branch of the
PICA (Amarenco and Hauw 1989). Although titles
241

reporting "cerebellar infarctions causing vertigo" are


correct on the basis of the imaging data and concurre nt clinical signs, it is more likely that the vertigo
itself results from pontomedullary brainstem
ischaemia ne ar the vestibular nuclei. The AICA supplies the rostral part of the vestibular nuclei, the
middle cerebellar peduncle, the flocculus, and the
neighbouring lobules of the cerebellum. The PICA
supplies the caudal part of the vestibular nuclei and
the dorsal medullary territory, the uvula, and the
nodulus (Amarenco and Hauw 1989). Partial occlusion of the AI CA can cause a lateral inferior pontine
infarction that mlmlCS vestibular neuritis
(Amarenco et al. 1990; Oas and Baloh 1992):
Occlusion of the vertebral artery or the PI CA causes
the typical Wallenberg's syndrome.
Most of the described cases have in common CT
scans that show cerebellar damage (both haemorrhagic and non-haemorrhagic infarctions).
Associated brainstem or cerebellar findings are central ocular motor signs, facial pain or numbness,
headache, or cerebellar ataxia. Peak incidence was in
the sixth decade, but occasionally individual patients
were under 30.
If caloric irrigation evokes bilateral normal
responses in a patient with rotational vertigo and
spontaneous nystagmus, vertigo of central origin
should be suspected. If, on the other hand, caloric irrigation elicits unilateral hyporesponsiveness and if the
vertigo is associated with ipsilateral hearing loss,
infarction of the AICA involving the internal auditory
and vestibular arteries should be suspected. Although
slow spontaneous recovery over a few weeks is the regular course in these small brainstem/cerebellar infarctions, a more severe course of progressive infarction
may develop after an interval of some hours or days
(Duncan et al. 1975; Sypert and Alvord 1975). Upbeat
or downbeat nystagmus syndrome (Chap. 11) indicates bilateral rather than unilateral brainstem dysfunction and may precede complete basilar artery
thrombosis. Such an experience led us to use anticoagulants more deliberately in cases of posterior fossa
ischaemia, in which the final outcome is unclear.

Vertigo

242

Hopf (1988) has described two patients with


symptoms and signs of vestibular neuritis, including
horizontal semicircular canal paresis. They also had
paresis of the masseter, temporal, and pterygoid
muscles as weIl as impairment of the masseter reflex
on the affected side. Hopf argued that infarction of a
small region ventrolateral to the floor of the fourth
ventricle was causative, since structures serving both
peripheral vestibular function and trigeminal motor
function are present in this area. This area is supplied by various branches of the basilar artery
(Hassler 1967; Duvernoy 1978). In most cases of central "pseudo" vestibular neuritis, additional signs
and symptoms help harden the diagnosis (p. 73).
Clinical examination and electrophysiological
determinations of brainstem reflexes help to detect
brainstem pathology as the cause of isolated cranial
nerve palsies (Hopf 1994; Thmke et al. 1995).
Findings are ocular motor abnormalities, pathological auditory evoked potentials (Wennmo and
Pyykk 1982; Ishikawa et al. 1993; Imate et al. 1995),
or central abnormalities of the masseter reflex (Hopf
1988) or the stapedius reflex (Bergenius and Borg
1983). Lacunar infarctions or demyelinating plaques
in multiple sclerosis at the root entry zone of the
eighth nerve can be demonstrated by MRI (Fig. 12.1;
Brandt et al. 1986; Francis et al. 1992). If vertigo and
lateropulsion are caused by an incomplete
Wallenberg's lateral medullary syndrome (p. 309),
the triad of Horner's syndrome, ipsilateral ataxia,
and contralateral hypalgesia is very helpful (Sacco et
al. 1993). The long-term prognosis of lateral
medullary infarctions is relatively good, and recurre nt posterior circulation strokes are uncommon
(Norrving and Cronqvist 1991).
The "epidemic central vestibular vertigo"
described by Pedersen (1959) and Merifield (1965) is
most likely a kind of benign viral encephalitis with
self-limited vertigo. Mangabeira-Albernaz and
Ganan<;:a (1988) presented a heterogeneous collection of eight patients, all of whom after different
viral infections complained of sudden prolonged
vertigo similar to that in vestibular neuritis, but
without horizontal semicircular canal paresis. The
patients had associated brainstem or cerebellar
deficits. Recovery was spontaneous, although probably
accelerated by corticosteroids, and occurred over
periods ranging from a few days to some months.

Paroxysmal central vertigo


Non-epileptic paroxysmal vertigo or other vestibular
syndromes may result from pathological excitation
of various vestibular structures (Brandt and

Dieterich 1994). Most of them occur in multiple


sclerosis, but others may be associated with a brainstern abscess (paroxysmal ocular tilt reaction,
Hedges and Hoyt 1982) or an arteriovenous malformation with previous bleeding (repetitive paroxysmal nystagmus and vertigo, Lawden et al. 1995).
Different manifestations of vestibular paroxysmia
of the brainstem have been described in multiple
sclerosis:

paroxysmal dysarthria, ataxia, and vertigo


(Andermann et al. 1959)
paroxysmal ocular tilt re action (Rabinovitch et
al. 1977) (p. 180)
paroxysmal room-tilt illusion (Dogulu and
Kansu 1997) (p. 224)

Paroxysmal dysarthria, vertigo and ataxia, as


described by Andermann et al. (1959), are weIl
known. They are probably the most common manifestations of paroxysmal attacks in multiple sclerosis
(Espir and Millac 1970). The mechanism of these
attacks has been suggested to be a transversally
spreading ephaptic activation of adjacent axons
within a partially demyelinated lesion in fibre tracts
of the pontine tegment um involving the brachium
conjunctivum (Osterman and Westerberg 1975). The
attacks may be the initial symptom of MS. They last
from a few seconds to a few minutes, and their frequency varies from a few per month to 200 per day.
Sometimes they are provoked by hyperventilation or
when rising. The character of the vertigo is rarely
rotatory but more typically a kind of missing postural
co ordination, with ataxia and broad-based gait often
associated with nystagmus or internuclear ophthalmoplegia.
Carbamazepine is a very effective treatment that
in most cases results in the complete disappearance
of the attacks (Espir and Millac 1970; Osterman and
Westerberg 1975) even in low dosages (200-400
mg/day). If untreated, they can continue for months
or years. If carbamazepine eliminates the paroxysms
for aperiod of weeks or months, the attempt should
be made to discontinue the medication. Most
patients will then remain free of paroxysms without
antiepileptics.
Familial periodic vertigo includes a variety of
autosomally dominant episodic ataxias (Griggs and
Nutt 1995; Brandt and Strupp 1997) (see Chap. 25),
which have recently been identified as channelopathies.
Paroxysmal vertigo evoked by lateral gaze. In the
absence of vestibular stimulation, prolonged lateral
eye deviation to the extreme left position for 5 to
20 s induced intense attacks of nystagmus and
rotational vertigo with postural imbalance lasting

Miscellaneous central vestibular disorders

from 50 to 90 s in a patient with repeated strokes in


the vertebrobasilar territory. Infarction mainly
affected the right dorsolateral medulla oblongata,
including the vestibular nuclei (Bttner et al. 1987).
During the attacks the nystagmus beat to the right
and counterclockwise, while the patient looked
around freely. A paroxysmal attack of nystagmus
and vertigo evoked by voluntary lateral eye position
is an unusual clinical finding. This distressing syndrome, which lasted for years, was refractory to
scopolamine. Attacks could also be elicited by rapid
head turns or tilts to one side. Thus, part of the
syndrome can be classified as a kind of central positional vertigo (p. 293). We speculated that the mechanism involved either canal-canal or canal-otolith
interaction, or a disturbance of the vestibular nuclei
gaze-holding mechanism (Bttner et al. 1987). To
date, we have seen two more cases with paroxysmal
vertigo evoked by sustained lateral gaze. This syndrome might therefore be less rare than was initially
assumed. As the procedure required to elicit such an
attack is unusual, similar cases might have been
overlooked. The patients have to look for several seconds (often more than 10 s) to the extreme left or
right before an attack occurs. An attack never
occurred if the eyes deflected 40 or less from middle
position. The resulting vertigo and discomfort are
such that the patients very sensibly avoid such
stimuli.
Attempted convergence may evoke congenital and
acquired forms of nystagmus (Sharpe et al. 1975;
Oliva and Rosenberg 1990) or change upbeat nystagmus to downbeat nystagmus (p. 206; Cox et al. 1981).
Sogg and Hoyt (1962) described a family with
attacks of vertical nystagmus, lasting from 20 to 90
minutes, which could be elicited by maintaining
upward gaze (Familial periodic ataxia; p. 365). These
attacks were, however, not accompanied by vertigo.
Thurston et al. (1985) described a patient with
periodically occurring conjugate eye and head turning, followed by nystagmus lasting for about 2 minutes which was associated with focal seizure activity
in the EEG (Vestibular epilepsy; p. 233).

Central vestibular falls without


vertigo
There are a few instances of what is probably central
vestibular dysfunction. In such cases, patients without paresis or sensory or cerebellar deficits are
unable to maintain an unsupported upright stance.
They do not, however, complain of vertigo. Their
conditions include thalamic astasia, lateropulsion in

243

Wallenberg's syndrome or in parieto-insular


vestibular cortex lesions (p. 192), and ocular-tilt
re action in pontomedullary or rostral midbrain
lesions (p. 312).
Thalamic astasia. Postural imbalance with a transient tendency to fall has been noted following therapeutic thalamotomy (Hassler 1966; Zoll 1969;
Velasco and Velasco 1979). It has been attributed to
muscle hypotonia or neglect and has also been
observed after thalamic infarctions (Cambier et al.
1980) and haemorrhages (Jenkyn et al. 1981; Verma
and Maheshwari 1986). Masdeu and Gorelick (1988)
described 15 patients with "thalamic astasia", in the
absence of motor weakness, sensory loss or cerebellar signs, due to lesions of different causes, all
primarily involving superoposterolateral portions of
the thalamus but sparing the rubral region.
"Typically, when asked to sit up, rather than using
the axial muscles, these patients would grasp the side
rail of the bed with the unaffected hand or with both
hands to pull themselves up." Thalamic astasia is
transient and lasts for days or weeks, with the dorsothalamic region being the criticallocus. Since posterolateral thalamic infarctions cause tilts of the
perceived vertical that are either ipsiversive or contraversive (Dieterich and Brandt 1993b), thalamic
astasia and tilts of perceived vertical may both
reflect a vestibular tone imbalance. Furthermore,
what Masdeu et al. (1994) described as astasia and
gait failure with damage of the pontomesencephalic
locomotor region involving the right pontine
peduncle area may be associated with vestibular dysfunction in roll. Their patient presented with a contraversive skew deviation of 10.
Thalamic hemiataxia differs from thalamic astasia
and rarely occurs in isolation; it is usually associated
with hemisensory loss without hemiparesis
(Solomon et al. 1994) or hemisensory loss and hemiparesis (Melo et al. 1992). The lesions involve the
ventral lateral nucleus of the thalamus and the adjacent posterior limb of the internal capsule and the
mid-to-posterior thalamus containing the dentatorubrothalamic and ascending pathways (Melo et
al. 1992; Solomon et al. 1994).
Lateropulsion
in
Wallenberg's syndrome.
Lateropulsion of the body is a well-known transient
feature of dorsolateral medullary infarction. These
patients have irresistible, ipsiversive falls but generally no subjective vertigo. Different brainstem
lesions from midbrain to medulla cause ipsiversive
deviation of the subjective vertical (Friedmann 1970;
Dieterich and Brandt 1993a). Transient ocular-tilt
re action (p. 179) and ipsiversive deviations of the
subjective vertical, which indicate a pathological
shift in the internal representation of the gravitational vector, are typically found in Wallenberg's

Vertigo

244

syndrome (p. 309) (Brandt and Dieterich 1987). We


hypothesised that the subjective vertigo is missing in
these patients (despite a striking tendency to fall
sideways), because individual muItisensory postural
regulation is adjusted to the deviated vertical.
Lateropulsion then represents postural compensation of an apparent body tiIt contraversive to the
lesioned side. Despite the resuIting postural imbalan ce and the conflicting true vertical, the body is
continuously adjusted toward what the central nervous system erroneously computes as vertical
(Brandt and Dieterich 1987). This could explain why
patients fall without vertigo or warning signals from
the muItisensory spatial orientation system.
Lateropulsion without hemiparesis also occurs in
corticallesions. These patients with infarctions of
the middle cerebral artery territory are weIl known
to physiotherapists, who call such patients "pushers".
It has been demonstrated that acute lesions of the
parieto-insular vestibular cortex cause contraversive
tiIts of the perceived visual vertical (Brandt et al.
1994), which makes it most likely that the cortical
lateropulsion is also due to a vestibular tone imbalance in the roll plane.
Drop attacks. Sheldon (1948, 1960) originally
described drop attacks, which also represent falls
without concurrent vertigo due to reflex-like hypotonia of antigravity muscles. The associated aetiology
is unknown in 69% of cases, cardiac in12%, cerebrovascular insufficiency in 8%, combined cardiac
and cerebrovascular disease in 7%, seizures in 5%,
psychogenic in 1%, and vestibular following late
Meniere's disease (Tumarkin's otolithic crisis, p. 94)
in only 3% of patients (Meissner et al. 1986). Patients
with drop attacks have a surprisingly good longterm prognosis. "Hyperekplexia", the sudden loss of
postural tone evoked by startling stimuli, was exacerbated by superimposed posterior thalamic infarction (Fariello et al. 1983).

Central vestibular syndromes in


multiple sclerosis
Following vascular aetiology, multiple sclerosis is
possibly the second most frequent cause of central
vestibular syndromes with vertigo, dizziness, and
postural imbalance. The paroxysmal vestibular syndromes are described separately (p. 242). The common occurrence of vertiginous symptoms in MS is
explained by the proximity of the vestibular nuclei
to the fourth ventricle (Bronstein and Rudge 1996);
plaques are common in periventricular areas. Not
only is the vestibular system frequently involved in

the course of MS (Grenman 1985), but at least 5% of


the patients have vertigo as the initial symptom
(McAlpine et al. 1972). An epidemiological study
found that the occurrence of vertigo at the onset of
MS is statistically a predictor of shorter life
expectancy (Riise et al. 1988). Vestibular syndromes
in MS are often associated with non-vestibular ocular motor disorders. MS plaques may mimic vestibular neuritis (p.241), cause paroxysmal vertigo,
ataxia, and disorientation (p. 242), upbeat/downbeat
nystagmus (p. 199), positional nystagmus, or positional vertigo (p. 291). Acquired ocular oscillations
that override fixation may cause distressing oscillopsia and postural imbalance: these include pendular
nystagmus, torsional more often than horizontal
(spontaneous) nystagmus, gaze-evoked nystagmus,
and internuclear ophthalmoplegia with dissociated
gaze-evoked nystagmus. The vestibular ocular reflex
may be disinhibited (impaired fixation suppression)
or inappropriate because of central vestibular or
ocular motor dysfunction, resulting in oscillopsia
and postural imbalance associated with head motion
or locomotion (p. 430).

Vestibular syndromes and brain


tumours
Hirose and Halmagyi (1996) state that balance disorders are common, while brain tumours are rare. An
isolated balance disorder is thus rarely the presenting symptom of a brain tumour. Nevertheless, some
patients with balance dis orders worry about having
a brain tumour, and some physicians also worry
about missing a brain tumour. The definite exclusion
of a brain tumour requires MRI and in some cases
brain biopsy. Supratentorial brain tumours rarely
cause vertigo, and if they do, they involve either the
temporoparietal or frontal regions. The vertigo is
described as dizziness with postural imbalance but
not as rotational vertigo. It is surprising that patients
presenting with tumours that affect the vestibular
cortex rarely complain of vertigo. If they report
vertiginous attacks, vestibular seizures should be
suspected (p. 234).
Infratentorial brain tumours causing vertigo
involve either the cerebellum (astrocytoma, ependymoma, medulloblastoma, haemangioblastoma,
metastasis) or the brainstem (glioma, cavernoma,
metastasis). Vestibular signs and symptoms depend
on the location: an ependymoma, arising on the floor
of the fourth ventricle will first affect pontomedullary vestibular nuclei and oculomotor structures (Frederickson and Fernandez 1964; Hirose and

245

Miscellaneous central vestibular disorders

Halmagyi 1996). A medulloblastoma arising on the


roof of the fourth ventride will first cause gait ataxia,
whereas tumours in the cerebellar hemispheres ean
become quite large without causing any detectable
vestibular dysfunetion. Initially brainstem tumours
often cause isolated cranial nerve palsies before
progressively affeeting long motor and sensory
pathways.

Metabolie disorders of the vestibular


system
Despite the eonsiderable literature on metabolie disorders eausing vestibular dysfunction (see review by
Rybak 1995), it remains obscure how diabetes mellitus,
hyperlipidaemia, hypothyroidism, and other hormonal disturbances can direcdy cause episodic or chronic vertigo. While all these eonditions have been
aceused, for example, of eausing Meniere's disease,
there is still no convincing argument explaining
either the meehanism that causes vertigo or why it
can be alleviated by diet or treatment with lipidlowering drugs.
The relationship between disturbances of the
autonomie nervous system and dizziness also
remains obscure. The only exception is dysautonomia,
in which postural changes and dis orders of the
autoregulatory meehanisms of cerebral blood fiow
can eause dizziness as the result of insufficient cerebral blood supply (Nakagawa et al. 1993).
Neuro-endoerinologieal factors in vertigo may
become of increasing interest as a picture of the
dose, but not as yet understood, relationships
between migraine and motion siekness (p. 334), vertigo, menstrual cydes, endocrine cydes, and emesis
slowly emerges (Gresty and Brookes 1997). As is
known in female migraine patients, some women
may experience heightened susceptibility to vertigo
and motion siekness at certain times during the
menstrual eyde.
Hypoactive vestibular responses to calorie and
rotational stimulation have been found in Wernicke's
encephalopathy (Furman and Becker 1989).
Superficial siderosis of the eentral nervous system
causing progressive hearing loss and intermittent
vertigo (Janss et al. 1993) can be explained by
peripheral vestibular involvement. It is unlikely that
abnormal responses to dynamic head and body tilts
in idiopathic Parkinson's disease refiect vestibular
dysfunetion (Bronstein et al. 1996; Paquet and
Hui-Chan 1997).

Referenees
Amarenco P, Hauw JJ (1989) Anatomie des arteres cerebelleuses
Rev Neurol (Paris) 145:267-276
Amarenco P, Roullet E, Chemouilli Ph, Marteau R (1990) Infarctus
pontin inferolateral: deux aspects cliniques. Rev Neurol (Paris)
146:433-437
Andermann F, Cosgrove JBR, Lloyd-Smith D, Walters AM (1959)
Paroxysmal dysarthria and ataxia in multiple sclerosis.
Neurology (Minneap) 9:211-215
Bergenius J, Borg E (1983) Audio-vestibular findings in patients
with vestibular neuritis. Acta Otolaryngol (Stockh) 96:389-395
Brandt Th, Dieterich M (1987) Pathological eye-head coordination in roll: Tonic ocular tilt reaction in mesencephalic and
medullary lesions. Brain 110:649-666
Brandt Th, Dieterich M (1994) Vestibular paroxysmia (Disabling
positional vertigo). Neuro-ophthalmology 14:359-369
Brandt Th, Dieterich M, Bchele W (1986) Postur al abnormalities
in central vestibular brainstem lesions. In: Bles W, Brandt Th
(eds) Disorders of posture and gait. Elsevier, Amsterdam, pp
141-156
Brandt Th, Dieterich M, Danek A (1994) Vestibular cortex lesions
affect the perception of verticality. Ann NeuroI35:403-412
Brandt Th, Strupp M (1997) Episodic ataxia type 1 and 2 (familial
periodic ataxia/vertigo). Audiol NeurootoI2:373-383
Bronstein AM, Rudge PC (1996) Vestibular dis orders due to multiple sclerosis, Arnold-Chiari malformation, and basal ganglia
dis orders. In: Baloh RW, Halmagyi GM (eds) Disorders of the
vestibular system. Oxford University Press, Oxford, pp 476-495
Bronstein AM, Yardley L, Moore AP, Cleeves L (1996) Visually and
posturally mediated tilt illusion in Parkinson's disease and in
labyrinthine defective subjects. Neurology 47:651-656
Bttner U, Straube A, Brandt Th (1987) Paroxysmal spontaneous
nystagmus and vertigo evoked by lateral eye position.
Neurology 37:1553-1555
Cambier J, Elghozi D, Strube E (1980) Usions du thalamus droit
avec syndrome de l'hemisphere mineur. Discussion du concept
de negligence thalamique. Rev Neurol (Paris) 136:105-116
Cox TA, Corbett JJ, Thompson HS, Lennarson L (1981) Upbeat
nystagmus changing to downbeat nystagmus with convergence.
Neurology 31:891-892
Dieterich M, Brandt Th (1993a) Ocular torsion and tilt of subjective visual vertical are sensitive brain signs. Ann Neurol
33:292 -299
Dieterich M, Brandt Th (1993b) Thalamic infarctions: differential
effects on vestibular function in the roll plane (35 patients).
Neurology 43: 1732-1740
Disher MJ, Telian SA, Kemink JL (1991) Evaluation of acute
vertigo: Unusuallesions imitating vestibular neuritis. Am J Otol
12:227-231
Dogulu CF, Kansu T (1997) Upside-down reversal of vision in
multiple sclerosis. J NeuroI244:461-472
Duncan GW, Parker SW, Fisher CM (1975) Acute cerebellar infarction in the PICA territory. Arch NeuroI32:364-368
Duvernoy AM (1978) Human brainstem vessels. Springer, Berlin
Heidelberg New York
Espir MLE, Millac P (1970) Treatment of paroxysmal disorders in
multiple sclerosis with carbamazepine (Tegretol). J Neurol
Neurosurg Psychiatry 33:528-531
Fariello RG, Schwartzman RJ, Beall SS (1983) Hyperekplexia exacerbated by occlusion of posterior thalamic arteries. Arch
NeuroI40:244-246
Francis DA, Bronstein AM, Rudge P, du Boulay EPGH (1992) The
site of brainstem lesions causing semicircular canal paresis: an
MRI study. J Neurol Neurosurg Psychiatry 55:446-449
Fredrickson JM, Fernandez C (1964) Vestibular dis orders in
fourth ventricle lesions. Arch OtolaryngoI80:521-540

246
Friedmann G (1970) The judgement ofthe visual vertical and horizontal with peripheral and central vestibular lesions. Brain
93:313-328
Furman JMR, Becker JT (1989) Vestibular responses in Wernicke's
encephalopathy. Ann NeuroI26:669-67 4
Grenman R (1985) Involvement of the audiovestibular system in
multiple sclerosis: an otoneurologic and audiologic study. Acta
Otolaryngol (Stockh) 420:1-95
Gresty M, Brookes G (1997) Deafness and vertigo. Curr Opin
NeuroI1O:36-42
Griggs RC, Nutt JG (1995) Episodic ataxias as channelopathies.
Ann NeuroI37:285-286
Guiang RL Jr, Ellington OB (1977) Acute pure vertiginous
dysequilibrium in cerebellar infarction. Eur NeuroI16:11-15
Hassler 0 (1967) Arterial pattern of human brainstem. Normal
appearance and deformation in expanding supratentorial conditions. Neurology 7:368-375
Hassler R (1966) Thalamic regulation of muscle tone and the
speed of movement. In: Purpura DP, Yahr MD (eds) The
thalamus. Columbia University Press, New York, pp 419-438
Hedges TR, Hoyt WF (1982) Ocular tilt reaction due to an upper
brainstem lesion: Paroxysmal skew deviation, torsion, and
oscillation ofthe eyes with head tilt.Ann Neurolll:537-540
Hirose G, Halmagyi GM (1996) Brain tumours and balance disorders. In: Baloh RW, Halmagyi GM (eds) Disorders of the
vestibular system. Oxford University Press, Oxford, pp 446-460
Hopf HC (1988) Vertigo and masseter paresis. A new local brainstern syndrome probably of vascular origin. J Neuro1235:42-45
Hopf HC (1994) Topodiagnostic value of brain stern reflexes.
Muscle Nerve 17:475-484
Imate Y, Sekitani T, Okami M, Miura M (1995) Central dis orders in
vestibular neuronitis. Acta Otolaryngol (Stoekh) Suppl
519:204-205
Ishikawa K, Edo M, Togawa K (1993) Clinical observation of 32
eases of vestibular neuronitis. Acta Otolaryngol (Stoekh) Suppl
503:13-15
Janss AJ, Galetta SL, Freese A, Raps EC, Curtis MT, Grossman RI,
Gomori JM, Duhaime AC (1993) Superficial siderosis of the
eentral nervous system: magnetie resonanee imaging and
pathological correlation. J Neurosurg 79:756-760
Jenkyn LR, Alberti AR, Peters JD (1981) Language dysfunetion,
somesthetie hemi-inattention, and thalamic hemorrhage in the
dominant hemisphere. Neurology 31: 1202-1203
Kmpf D (1986) Der benigne pseudovestibulre Kleinhirninsult.
Nervenarzt 57:163-166
Lawden MC, Bronstein AM, Kennard C (1995) Repetitive paroxysmal nystagmus and vertigo. Neurology 45:276-280
Mangabeira-Albernaz PL, Ganan~a MM (1988) Sudden vertigo of
eentral origin. Aeta Otolaryngol (Stoekh) 105:564-569
Masdeu JC, Goreliek PB (1988) Thalamie astasia: Inability to stand
after unilateral thalamie lesions. Ann Neuro123:596-603
Masdeu JC,Alampur U, Cavaliere R, Tavoulareas G (1994) Astasia
and gait failure with damage of the pontomeseneephalie loeomotor region. Ann Neuro135:619-621
MeAlpine D, Lumsden CE, Aeheson ED (1972) Multiple sclerosis: a
reappraisal. Churehill Livingstone, Edinburgh
Meissner I, Wiebers DO, Swanson JW, O'Fallon WM (1986) The
natural history of drop attacks. Neurology 36:1029-1034
Melo TP, Bogousslavsky J, Moulin T, Nader J, Regli F (1992)
Thalamie ataxia. J Neuro1239:331-337
Merifield DO (1965) Self-limited idiopathie vertigo (epidemic vertigo). Areh OtolaryngoI81:355-358

Vertigo
Nakagawa H, Ohashi N, Kanda K, Watanabe Y (1993) Autonomie
nervous system disturbance as aetiologieal baekground of vertigo and dizziness. Aeta Otolaryngol (Stoekh) Suppl
504:130-133
Norrving B, Cronqvist S (1991) Lateral medullary infaretion:
prognosis in an unseleeted series. Neurology 41:244-248
Oas JG, Baloh RW (1992) Vertigo and the anterior inferior eerebellar artery syndrome. Neurology 42:2274-2279
Oliva A, Rosenberg ML (1990) Convergenee-evoked nystagmus.
Neurology 40:161-162
Osterman PO, Westerberg CE (1975) Paroxysmal attaeks in multiple sclerosis. Brain 98: 189-202
Paquet N, Hui-Chan CWY (1997) Responses to dynamie headand-body tilts are enhaneed in Parkinson's disease. Can J
Neurol Sei 24:44-52
Pedersen E (1959) Epidemie vertigo. Clinieal picture, epidemiology,
and relation to eneephalitis. Brain 82:566-580
Rabinoviteh HE, Sharpe JA, Sylvester TO (1977) The oeular tilt
reaetion. A paroxysmal dyskinesia associated with elliptieal
nystagmus. Areh OphthalmoI95:1395-1398
Riise T, Bronning M, Aarli JA, Nyland H, Larsen JP, Edland A (1988)
Prognostie faetors for life expeetaney in multiple sclerosis
analysed by Cox models. J Clin Epidemio141: 1031-1036
Rubinstein RL, Norman DM, Schindler RA, Kaseff L (1980)
Cerebellar infaretion. A presentation of vertigo. Laryngoseope
90:505-514
Rybak LP (1995) Metabolie disorders of the vestibular system.
Otolaryngol Head Neek Surg 112:128-132
Saceo RL, Freddo L, Bello JA, Odel JG, Onesti ST, Mohr JP (1993)
Wallenberg's lateral medullary syndrome. Areh Neurol
50:609-614
Samson M, Mithout D, Thiebot J, Segond D, Weber J, Proust B
(1981) Forme benigne des infaretus eerebelleux. Rev Neurol
(Paris) 137:373-382
Sharpe JA, Hoyt WF, Rosenberg MA (1975) Convergence-evoked
nystagmus. Congenital and aequired forms. Areh Neurol 32: 191
Sheldon JH (1948) The soeial medieine of old age: report of an
enquiry in Wolverhampton. Oxford University Press, London
Sheldon JH (1960) On the natural history of falls in old age. Br
Med J (Clin Res) 2:1685-1690
Sogg RL, Hoyt WF (1962) Intermittent vertieal nystagmus in a
father and son. Areh Ophthalmol 68:515-517
Solomon DH, Barohn RJ, Bazan C, Grissom J (1994) The thalamie
ataxia syndrome. Neurology 44:810-814
Sypert GW, Alvord EC (1975) Cerebellar infaretion. A clinieopathologie study. Arch Neuro132: 357-363
Thmke F, Tettenborn B, Hopf HC (1995) Third nerve palsy as the
sole manifestation of midbrain isehemia. Neuro-ophthalmology
15:327-335
Thurston STE, Leigh RJ, Osoria I (1985) Epileptie gaze deviation
and nystagmus. Neurology 35: 1518-1521
Velaseo F, Velaseo M (1979) A retieulothalamie system mediating
proprioeeptive attention and tremor in man. Neurosurgery
4:30-36
Verma AK, Maheshwari MC (1986) Hyperesthetie-ataxiehemiparesis in thalamie hemorrhage. Stroke 17:49-51
Wennmo C, Pyykk I (1982) Vestibular neuritis. A clinical and
eleetro-oeulographie analysis. Acta Otolaryngol (Stoekh)
94:507-515
Zoll JG (1969) Transient anosognosia assoeiated with thalamotomy;
is it eaused by proprioeeptive loss? Confin Neuro131:48-55

SECTION D
Positional and positioning vertigo

Positional or positioning vertigo and nystagmus


syndromes can be attributed to either peripheral or
central vestibular dysfunction (Table D.l). The term
positional signifies that an altered otolith input,
which
results
from
a
newly
assumed
head pOSltlOn, is the causative mechanism.
"Positioning" refers to the change in head position
itself (head movement relative to the gravity vector)
as the causative mechanism. The most common
form is benign paroxysmal positioning vertigo
(Chap. 16). It is due to canalolithiasis in the posterior
semicircular canal (p-BPPV) or less frequently in the
horizontal (h-BPPV) or anterior (a-BPPV) semicircular canal. Positional vertigo is also a characteristic
symptom of perilymph fistulas (Chap. 6), Meniere's
disease (Chap. 5), and vestibular atelectasis.
Other labyrinthine manifestations such as positional alcohol nystagmus (p. 286), positional nystagmus with macroglobulinaemia (p. 288) and "heavy
water" or glycerol ingestion (p. 287) occur because of
a specific gravity differential between the cupula and
the endolymph (buoyancy mechanism, Chap. 17).
Neurovascular cross-compression of the vestibular
nerve is the causative factor for vestibular paroxysmia ("disabling positional vertigo"), an insufficiently
described entity (Chap. 17). Head-position dependent eighth nerve compression with positional vertigo mayaIso be due to a mass in the cerebellopontine
angle (p. 124).
Central positional vertigo (Chap. 18) is induced
either by head movements, wh ich result in a transient ischaemia of the pontomedullary brainstem
(transient vertebrobasilar ischaemic attacks), or
more commonly in archicerebellar lesions by a
change in head position relative to the gravitational
vector. The latter can take at least three forms: positional downbeat nystagmus (nodulus), positional

nystagmus without concurrent vertigo and paroxysmal positional vertigo with nystagmus and/or
nausea (vomiting). Central vestibular positional
vertigo syndromes always indicate a dysfunction
(disinhibition of otolith-canal interaction) of the
infratentorial connections between the vestibular
nuclei and the intra-axial vestibulocerebellar, in particular dorsal vermis, structures.
Table 0.1.

Positional or positioning vertigo and nystagmus

Central vestibular (vestibular nuclei, dorsal vermis, vestibulocerebellar

pathways)
Positional downbeat nystagmus
Central positional nystagmus without major vertigo
Central paroxysmal positional vertigo with nystagmus and/or nausea
Transient, head-position dependent vertebrobasilar ischaemia

Vestibular nerve

Neurovascular cross-compression ("disabling positional vertigo":


"vestibular paroxysmia")
Head-position dependent eighth nerve compression with transient
vestibular loss

Peripherallabyrinth

Benign paroxysmal positioning vertigo (BPPV)


- Of the posterior semicircular canal (p-BPPV)
- Of the horizontal semicircular canal (h-BPPV)
- Of the anterior semicircular canal (a-BPPV)
Cupula/endolymph specific -gravity differential (buoyancy mechanism)
- Positional alcohol vertigo/nystagmus (PAN)
- Positional "heavy water" nystagmus
- Positional glycerol nystagmus
- Positional nystagmus with macroglobulinaemia
Perilymph fistula
Meniere's disease
Vestibular atelectasis

Physiological vertigo

- "Head (neck) extension vertigo"


- "Bending-over vertigo"

Psychogenic positionallpositioning vertigo

249

Benign paroxysmal positioning vertigo

Benign paroxysmal positioning vertigo (BPPV; also


known as positional vertigo) was initially defined by
Baniny in 1921. The term itself was coined by Dix
and Hallpike (1952). Lanska and Remler (1997)
describe in detail the history of BPPV, its original
description, the proper eponymic designation for the
provocative positioning test, and the steps leading to
our current understanding of its pathophysiology.
BPPV is the most common cause of vertigo, partkularly in the elderly. By age 70, about 30% of al1 elderly
subjects have experienced BPPV at least once. This
condition is characterised by brief attacks of rotatory
vertigo and concomitant positioning rota tory-linear
nystagmus which are elicited by rapid changes in
head position relative to gravity. BPPV is a mechankai dis order of the inner ear in whkh the precipitating positioning of the head causes an abnormal
stimulation, usually of the posterior semicircular
canal (p-BPPV) of the undermost ear, less frequently
of the horizontal semicircular canal (h-BPPV).
Schuknecht (Schuknecht 1969; Schuknecht and
Ruby 1973) hypothesised that heavy debris settle on
the cupula (cupulolithiasis) of the canal, transforming it from a transducer of angular acceleration into
a transducer of linear acceleration.lt is now gene rally
accepted that the debris float freely within the
endolymph of the canal ("canalolithiasis") (Parnes
and McClure 1991; Epley 1992; Brandt and Steddin
1992). The debris - possibly partieIes detached from
the otoliths - congeal to form a free-floating eIot
(plug). Since the elot is heavier than the endolymph,
it will always gravitate to the most dependent part of
the canal du ring changes in head position which
alter the angle of the cupular plane relative to gravity.
Analogous to a plunger, the elot induces bidirection
al (push or pull) forces on the cupula, thereby triggering the BPPV attack. Canalolithiasis explains a11
the features of BPPV: latency, short duration, fatiguability (diminution with repeated positioning),
changes in direction of nystagmus with changes in
head position, and the efficacy of physical therapy
(Brandt and teddin 1993; Baloh et al. 1993; Brandt
et a1. 1994).

In 1980 Brandt and Daroff proposed the first


effective physkal therapy (positioning exercises) for
BPPV. Based on the assumption that cupulolithiasis
was the underlying mechanism, the exercises were a
sequence of rapid lateral head/trunk tilts, repeated
serially to promote loosening and, ultimately, dispersion of the debris toward the utricular cavity. In 1988
Semont and co-workers introduced a single liberatory
manoeuvre, and Epley promoted a variation in 1992,
which Herdman et al. (1993) later modified. If performed properly, a11 three form of therapy (BrandtDaroff exercises and Semont and Epley's liberatory
manoeuvres) are effective in BPPV patients
(Herdman 1990; Herdman et al. 1993). The efficacy
of physkal therapy makes selective surgical destructions such as transsection of the posterior nerve
(Gacek 1978) or non-ampullary plugging of the posterior semicircular canal (Pace-Balzan and Rutka
1991) largely unnecessary.
About 5-10% ofBPPV patients suffer from horizontal canalolithiasis (h-BPPV; McClure 1985).
h-BPPV is elicited when the head of the supine
patient is turned from side to side, around the longitudinal z-axis. Combinations are possible, and transitions from p-BPPV to h-BPPV occur, if the elot
moves from one to the other semicircular canal.
Transitions from canalolithiasis to cupulolithiasis in
h-BPPV patients have been described (Steddin and
Brandt 1996). Most of the cases appear to be idiopathic (degenerative?), their incidence increasing
with advancing age. Prolonged bedrest also facilitates their occurrence. Other cases arise due to
trauma, vestibular neuritis, or inner ear infections.
The diagnosis of typical BPPV is simple and safe:
the patient must have the usual his tory and exhibit
positioning nystagmus toward the causative, undermost ear. Diagnosis is less easy in rare ca ses, for
example, in patients with horizontal semicircular
canal cupulolithiasis (p. 270) who exhibit positional
nystagmus beating toward the uppermost ear for
several minutes. Differential diagnosis ineludes different forms of central vestibular vertigo or nystagmus, vestibular paroxysmia, perilymph fistula, drug

251

252

Vertigo

or alcohol intoxication, vertebroba ilar ischaemia,


Meniere's disease, and psychogenic vertigo.

The clinical syndrome

Patients with typical BPPV report attacks of rotatory


vertigo, postural imbalance, and sometimes nausea
precipitated by the following manoeuvres:

sitting up from a supine position (particularly


after awaking in the morning)
when first lying down in bed
turning over in bed from one side to the other
extending the neck (head) to look up or get
something from above
flexing the neck (head) when bending over

The occurrence of BPPV in the supine position is


very disturbing and makes patients afraid of falling
backward, an almost unique complaint. In the
upright position, vertigo attacks produced by
changes in head position are incapacitating and can
be dangerous, for example when a sufferer is looking
up at the ceiling while standing on a ladder. In such a
situation BPPV can cause a catastrophic fall.
Sometimes the "probable" diagnosis of BPPV is
simply based on the typical patient history, because
the condition has spontaneously resolved by the
time of examination in accordance with its usually
benign course. However, there is no absolute reliability of a diagnosis based on history, and some
patients may describe their vertigo in a rather atypical way (Norre 1995).
Most patients are aware that tilting or rotating of
the head toward the right, the left, or in both directions will induce the attacks. Since as a rule the first
positioning manoeuvre triggers the strongest attack
of vertigo and the most obvious positioning nystagmus, the first examination of a patient with a his tory
of BPPV should always be a positioning manoeuvre
toward the side of the posterior semicircular canal
thought to be affected. This initial step is all the
more important as repeated positioning manoeuvres
cause fatigue of the induced vertigo and nystagmus,
thus making diagnostic evaluations more difficult
and uncertain.
The following rules have proven sound for examining patients with BPPV:

Positioning testing should be done first during


the physical examination, and the initial positioning manoeuvre should be directed toward
the ear assumed to be affected, because vertigo

and nystagmus will fatigue during repeated


manoeuvres.
Frenzel's lenses should be used whenever possible,
to avoid partial suppression of the positioning
nystagmus by fixation.
Patients should be instructed before the positioning manoeuvre to ensure better cooperation.
They should also be asked not to shut their eyes
when vertigo begins.
Vertigo and nystagmus are maximal if the
patient is positioned rapidly and the head is
abruptly halted at its final position.

We usually perform the lateral head-trunk tilt with


the patient in a sitting position on a couch (Figs. 16.1
and 16.2). Others prefer the so-called Dix-Hallpike
manoeuvre (1952), during wh ich the patient is tilted
backward until the head is both turned and hanging
(Fig.2.16).
To confirm a suspected canalolithiasis of the posterior semicircular canal of the right labyrinth, the

iJL

(
Fig.16.1. Benign paroxysmal positioning vertigo and nystagmus are precipitated by rapid lateral head tilt toward the affected
ear or by neck extension (top). The typical nystagmus (best seen
with Frenzel's glasses) beats toward the undermost ear (or
upward), rotating counterclockwise with right ear lesions (bottom left),and clockwise with left ear lesions.The rotatory-linear
nystagmus reflects ampullofugal stimulation of the posterior
semicircular canal with activation of the ipsilateral superior
oblique and contra lateral inferior rectus eye muscles (bottom
right).

253

Benign paroxysmal positioning vertigo

head of the sitting patient is rotated 45 to the left.


Then the observer tilts the passive patient quickly to
the right side. After a latency of seconds rotatory
vertigo and positioning nystagmus occur in a
crescendo/decrescendo mode. The horizontal rotatory

nystagmus beats toward the undermost ear. After


vertigo and nystagmus cease, the patient is again
quickly returned to the initial upright position. In
most cases this will result in a less violent and shorter vertigo and nystagmus toward the opposite
direction.
To confirm a suspected canalolithiasis of the posterior semicircular canal of the left labyrinth, the
head of the sitting patient is turned 45 to the right
and the tilting manoeuvre is performed to the left
for the first time. Thereafter one should always check
for canalolithiasis of the horizontal semicircular
canal (p. 269), since combinations ofboth conditions
may manifest simultaneously in the same patient.
Finally, one should check for rare anterior canal
BPPV (p. 279).
Observation of the positioning nystagmus provides the definitive diagnostic criteria for typical
p-BPPV. They include

latency: vertigo and nystagmus begin one or


more seconds after the head is tilted toward the
affected ear and increase in severity to a maximum.
duration less than 40 s: nystagmus gradually
reduces after 10-40 sand ultimately abates even
when the precipitating head position is maintained.
linear-rotatory nystagmus: the nystagmus is best
seen with Frenzel's glasses (for example, lenses
+ 16 dpt), which prevents suppression by fixation.
The nystagmus is linear-rotatory, with the fast
phase beating toward the undermost ear or
upward when gaze is directed to the uppermost
ear.
reversal: when the patient returns to the seated
position, the vertigo and nystagmus may reoccur
less violently in the opposite direction.
fatiguability: constant repetition of this man oeuvre
will result in ever-lessening symptoms.

These five criteria are crucial for further discussion


of the confusing literature on the mechanism of
BPPV. They provide the major arguments to
prove or disprove any hypothetical explanation of
cupulolithiasis or canalolithiasis as the causative
factor (p. 259).

Positioning nystagmus

Fig. 16.2. Precipitating positioning manoeuvres for typical


posterior semicircular canal BPPV: lateral head-trunk tilt toward
the affected ear.

The direction of paroxysmal positioning nystagmus


(best observed when the patient wears Frenzel's
glasses) is similar in all cases of p-BPPV: the nystagmus beats toward the undermost affected ear,
and has a rotatory component either clockwise
(when following precipitate leftward movement) or

254

counterdockwise (when following precipitate rightward movement) from the viewpoint of the observer. This pattern of eye movements and the
characteristics of a short latency, limited duration,
reversal on returning to the upright position, and
fatiguability on repetitive provocation are typical.
A doser look, however, in particular at the gazedependent differential effects on the direction and
the conjugation of induced eye movements, reveals
much more complexity and also explains some of
the seemingly contradictory descriptions in the literature (Fig. 16.3). What Harbert rediscovered in 1970
was already part of Baniny's original description in

Fig.16.3. In posterior canal BPPV the nystagmus is linear-rotatory, with the fast phase beating toward the undermost ear (top)
or upward toward the forehead, when the gaze is directed
toward the uppermost ear (bottom). Ampullofugal stimulation of
the posterior semicircular canal causes excitation of the ipsilateral superior oblique and the contra lateral inferior rectus muscles,
which drive both eyes to move downward with the slow phases
(bottom, filled arrows) and upward with the quick phases of the
nystagmus (bottom, open arrows). (hanges in the beating direction as weil as minor disconjugation of the eyes can be explained
by the different angle of insertion of the oblique and rectus
muscles.

Vertigo

1921. Positional nystagmus is mainly rotatory and


beats toward the undermost ear when gaze is directed
toward the undermost ear. However, it beats mainly
in a linear-oblique direction toward the forehead
when the gaze is directed toward the uppermost ear.
The neuronal network mediating the vestibuloocular reflex of the vertical canals is based upon sensory convergence within a three-neuron reflex are. It
links a set of extraocular musdes with their primary
action aligned to the particular spatial plane of either
the anterior or posterior canal. Sensorimotor transformation from canal planes to the planes of eye
movements has been demonstrated at the level of second-order neurons within the vestibular nudei. These
neurons project to oculomotor neuron pools and
always contact their two respective principal eye
movers in the excitatory as weH as in the inhibitory
(push-puH operational mode) vestibulo-ocular motor
link (Graf et al. 1983; Graf and Ezure 1986).
AmpuHofugal stimulation of the posterior semicircular canal causes excitation of the ipsilateral
superior oblique and the contralateral inferior rectus
musde (Cohen et al. 1966; Graf et al. 1983). This
causes both eyes to move downward with the slow
phases and to move upward with the quick phases of
the nystagmus (Fig. 16.3). Monocular ENG re cordings of horizontal and vertical components demonstrated a larger horizontal component in the
ipsilateral eye and a larger vertical component in the
contralateral eye, both of which can be explained by
the different angle of insertion of the oblique and
rectus musdes (Baloh et al. 1979). Furthermore, the
amplitude of the horizontal component in each eye
depends on the direction of gaze relative to the head.
ENG recordings sometimes appear inconsistent on
visual examination, thus leading to reports of reversals of the horizontal component with changes in
gaze position and dissociated eye movements as weH
as downbeating nystagmus (Katsarkas and
Outerbridge 1983; Katsarkas 1991; Baloh et al. 1987).
Whereas Yagi and Ushio (1995) met with difficulties
using three-dimensional eye-movement analysis to
support the view that typical BPPV originates from
the posterior canal, Fetter and Sievering (1995) took
a different approach. They used search coils to
measure 3-D eye position and were able to relate the
nystagmus to the particular posterior canal as
expected from dinical observation.

Vertigo and posture


While the nystagmus pattern in BPPV attacks has
been weH described, less information is available on
the direction and magnitude of postural destabilisa-

255

Benign paroxysmal positioning vertigo

tion. Posturographic measurements have been made


in patients with BPPV, in whom attacks were elicited
by head tiIt while they were standing on a forcemeasuring posturographic platform. These measurements revealed a characteristic pattern of postural
instability (Bchele and Brandt 1979): after a short
latency patients exhibit large sway amplitudes,
predominantly in the fore-aft direction (Figs.
16.4-16.6), with a mean sway frequency range below
3 Hz. The amount of instability decreases gradually
over 30 seconds in parallel to the diminishing sensation of vertigo and nystagmus. When subjects close
their eyes, the acute destabilisation may lead to an
almost irresistible tendency to fall, which can only
be overcome by the intervention of corrective
somatosensory input. Head tilts without typical vertigo attacks resuIt only in a discrete destabilisation
due to the slight shift of body mass. Patients in
whom an attack could be elicited while they were
standing on a force-measuring platform show a shift
of the mean position of the cent re of gravity forward

f}

t Nm

, Nm' }
I

f
fore - aU
lateral

9.25 (righ t)

head lilt

~-

no vertigo

____

~_~

_ _ __

-----.

5 s

d'.49 (leU)

in the direction of the head tilt (Figs. 16.4 and 16.6),


with a concurrent increase in sway amplitude. Body
displacement forward subsides within about 30 s.
Mean sway amplitudes increase by a factor of 4 in
the fore-aft plane and by a factor of 3 in the lateral
plane. Some patients exhibit a superimposed 3-Hz
fore-aft sway (Fig. 16.5); an oscillator-like tremor of
body posture has been described in chronic alcoholics
with anterior lobe cerebellar atrophy (Dichgans et al.
1975). This 3-Hz tremor is sometimes concluded to
be pathognomonic for anterior lobe cerebellar atrophy. However, since none of the patients with BPPV
who were tested, suffered from cerebellar disease,
such a conclusion is not justified. The measurable
shift of the cent re of gravity in the forward direction
and ipsilateral to the tiIted head (Fig. 16.6) can be
interpreted to be the motoric compensation for the
initial subjective vertigo in the opposite direction,
the diagonal plane being coplanar to the spatial
plane and working range of the ipsilateral posterior
canal (Bchele and Brandt 1986; Brandt and
Dieterich 1993). This compensation for the initial
"subjective fall" caused by the lesion occurs in such a
way that a "measurable fall" occurs in the opposite
direction. Posturographic results are thus consistent
with posterior canal stimulation.
Black and Nashner (1984) claimed to distinguish
three groups of patients by posturographic measurements: those with unilateral or bilateral loss of
peripheral vestibular function, those with BPPV
without peripheral vestibular deficit, and those with

head upright (eyes closed)

, nead ti lt
tore - afl
t-------<

d'. 58

paro.y5m31 positioning vertigo (leU)

55

head uprighl (eye5 closed )

9.55 (roght )

4
t Nm

LO::"--..........->< _______ __________ _ _ -- - -

d.

- --- - - - - --

Fig.16.4. Original recordings of fore-aft and lateral body sway


offour patients suffering from p-BPPV. When vertigo is elicited
(patients 2-4), a destabilisation mainly in the fore-aft direction
occurs with simultaneous increase in the amplitude of sway and
a shift of the centre of gravity forward and laterally; the lateral
shift always occurs toward the affected ear. If no vertigo is
evoked, postural instability is minimal and reflects only the slight
shift of body mass due to head tilt (Patient 1). (From Bchele and
Brandt 1979.)

1 Nm

57 anterior lobe atrophy (alcohollc)

0------<

9. 37

acut e hydanloin intox ica ti on

Fig.16.5. Examples of occasionally superimposed 3-Hz fore-aft


body sway in a patient with p-BPPV during an attack (top) and
the oscillator-like 3-Hz body tremor in a patient with (alcoholic)
late anterior lobe cerebellar atrophy (middle) and in a patient suffering from acute hydantoin intoxication (bottom).

256

Vertigo
head Hit
fo re-alt sway

lateral sway

lN:l-~~~---
I
~
o

10

20

30

Fig.16.6. Mean amplitudes of body sway with an approximation of the deviation of the centre of gravity in 13 patients with pBPPV, in whom an attack could be elicited. The body shifts
forward and ipsilateral to the affected ear, and increased sway
amplitudes decrease along with a lessening of nystagmus and
vertigo. (Bchele and Brandt 1979.)

a combination of both BPPV and a peripheral


vestibular deficit. Patients were exposed to either a
stable or a moving foot support, and to variable visual
surroundings (Black et al. 1983). The authors concluded from their results that the BPPV patients, in
the main, use visual information to compensate for
the postural destabilisation induced by a vestibular
irritation, whereas a peripheral vestibular deficit
causes a disturbed adaptive reorganisation of visual
and postural references for orientation in space.
Vestibulospinal compensation results, therefore,
from either the suppression of vestibular inputs or
the simultaneous selection of an alternative reference for orientation: i.e. either vision or support.
This view, however, was not confirmed by Voorhees

Anastasopoulus et al. (1997) tried to assess utricular dysfunction in 14 patients with BPPV by linear
VOR during lateral whole body translation. They
explained the absence of gross changes of the lateral
VOR as due to either functionaIly irrelevant utricular damage or central compensation of a chronic
deficit. It is not easy to examine utricular function in
isolation (Fluur 1974) in human subjects, but one
possible source of information is dynamic ocular
counterrolling, a reflex that apparently depends on
the utricular otolith organs. Markharn and Diamond
(Markharn et al. 1987) described abnormal ocular
counterroIling in 10 of 18 patients tested by rotating
the body about two axes with the head fixed in relation to the body. The two axes were the naso-occipital
and the submental-vertex (barbecue rotation); the
latter provided a more sensitive indicator of utricular dysfunction. The most common dysfunctions
were disconjugate eye movements and hypoactivity.
Subjects were more sensitive to ipsilateral than to
contralateral tilt. This agrees with our own experience in patients with acute BPPV, some of whom
show significant deviations of the subjective visual
vertical at the initial stage when they adjust a test bar
to the perceived vertical.

Natural course
The natural history of BPPV is considered benign
because it resolves spontaneously within weeks or
months in most patients. However, in about 20-30%
of the patients the condition persists when untreated,
and it recurs in another 30% after variable periods
for years.

Differential diagnosis

(1989).

Clinicians are weIl aware that patients in the acute


phase of BPPV (and sometimes foIlowing a successful liberatory manoeuvre) also complain of
unsteadiness in gait and postural balance, which
they describe as "walking on pillows". These symptoms can be classified as otolithic vertigo. They can
be attributed to a suddenly unequal weight of the
two utricles on the maculae, which generates a
vestibular tone imbalance. This is above aIl true in
post-traumatic BPPV patients, in whom there is an
acute loosening of otoconia from the macula bed on
one or both sides. There is usuaIly a gradual
improvement in the control of posture and gait in
the days or weeks foIlowing the onset of the illness:
this reflects both the process of central compensation and, more slowly, peripheral restitution of the
equal weights.

The most important differential diagnosis is that of


central vestibular positional nystagmus (p. 291). It
should be suspected in aIl cases of positional nystagmus without concomitant subjective vertigo,
although some lesions around the fourth ventricle
mayaiso induce very violent vertigo and nausea.
Central positional nystagmus does not subside with
maintenance of the head in the precipitating position, it may not exhibit fatiguability with repetitive
stimulation, it may change its direction when different head positions are assumed, or it may occur as
downbeat nystagmus only in the head-hanging position. Frequently, additional neurological (in particular oculomotor) signs point toward a central cause of
the symptoms.
BPPV must also be differentiated from positional
nystagmus in Meniere's disease (p. 83), perilymph

257

Benign paroxysmal positioning vertigo

fistulas (p. 99), drug and/or alcohol intoxication


(p. 286), neurovascular cross-compression (vestibular paroxysmia), and rare conditions such as
Waldenstrm's macroglobulinaemia.

Pathomechanism and aetiology


Pathomechanism
Typical posterior canal BPPV is caused by
canalolithiasis, a free-floating clot within the
endolymph of the posterior canal. A clot-induced
endolymph flow mechanism is compatible with all
features of BPPV such as latency, limited duration,
fatiguability, change in direction of the induced nystagmus, and the efficacy of physical therapy for both
posterior and horizontal semicircular canal BPPV.
There was, however, a long-continuing controversy
in the quite extensive literature, which first arose
over the following questions:

Is BPPV a peripheral or a central vestibular dysfunction?


Is cupulolithiasis compatible with the clinical
features of BPPV?
Is canalolithiasis compatible with the clinical
features of BPPV?

In the following, these three questions will be discussed on the basis of the contradictory literature.
Peripheral or central vestibular dysfunction?

There is convincing evidence for a peripheral origin


of BPPV as summarised in Table 16.1. The critical
role of the posterior semicircular canal in BPPV is
supported by the results of ampullofugal cupula
deflection: the ipsilateral superior oblique and the
contralateral inferior rectus eye muscles contract
(Szentagothai 1950). These muscles must be activated
for the typical nystagmus to be induced. The importance of the posterior semicircular canal is also consistent with the pattern of eye movements produced
by its selective stimulation, which has been dem onstrated in the monkey (Cohen et al. 1966).
Posturographic measurements in patients with pBPPV have shown a forward shift of the centre of
gravity ipsilateral to the tilted head, which is coplanar with the spatial plane and working range of the
ipsilateral posterior canal (Bchele and Brandt
1986). This is further supported by findings of Gacek
(1984), who cured patients with chronic unilateral
BPPV by selective dissection of the ipsilateral posterior

Table 16.1. Arguments for BPPV as a peripherallabyrinthine disorder


Common association of BPPV with disorders of the labyrinth:
Head trauma, post- stapes surgery
Vestibular neuritis
Chronic suppurative otitis media
Degenerative changes of ageing
Occlusion of anterior vestibular artery
Success of labyrinthine and vestibular nerve surgery:
Labyrinthectomy
Vestibular neurectomy
Dissection of posterior ampullary nerve
Canal plugging
Histological findings:
Debris in semicircular canals
Debris fixed on the cupula
Degenerated utricular macula(?)
Positioning nystagmus compatible with semicircular canal stimulation:
Ampullofugal stimulation of posterior semicircular canal = p-BPPV
Ampullopetal stimulation of horizontal semicircular canal = h-BPPV
Spontaneous course with unpredictable fatigue and remission:
Varying severity and provocability of BPPV attacks with repetitive
positioning manoeuvres
Absence of clinical and ENG signs of central involvement
(From Brandt and Steddin 1993.)

ampullary nerve. Parnes and McClure (1990, 1991)


and Pace-Balzan and Rutka (1991) also reported that
patients with intractable BPPV were cured by transmastoid posterior semicircular canal occlusion.
The clinical relevance of central "pseudo BPPV"
has been repeatedly stressed (Gregorius et al. 1976;
Watson et al. 1981; Sakata et al. 1987; 1991). However,
it should not be overestimated for it is quite rare.
Even large series of several hundreds of patients
with BPPV do not support central vestibular pathology (Baloh et al. 1987; Katsarkas 1991). A variety of
central positional vertigo syndromes occur with a
mass (tumour, haematoma) near the fourth ventricle
and the vestibular nuclei (Brandt 1991). Evidence
from animal experiments shows that posterior cerebellar vermis lesions may cause positional nystagmus, which mimics BPPV (Allen and Fernandez
1960).
The tradition al view of cupulolithiasis

The early assumption by Barany (1921; Dix and


Hallpike 1952) that the underlying lesion must be
situated in the vestibular end-organ and must
involve the otolith was later supported by
Schuknecht and Ruby (Schuknecht 1962, 1969;
Schuknecht and Ruby 1973), who postulated a
mechanical pathogenesis, which they termed "cupulolithiasis". They found basophilic deposits on the
cupula of the causative posterior semicircular canal
in individual patients who manifested unilateral
BPPV prior to death from unrelated disease. These

258

Fig. 16.7. Schematic drawing of the left vestibular labyrinth as


seen from the occiput.lnset shows the cu pu la of the posterior
semicircular canal with detached otoconial sediment on both
sides of the cu pu la and in the posterior canal. (From Brandt and
Steddin 1993.)

deposits exceeded the size of those found in more


than 30% of temporal bones in a control population.
They argued that inorganic particles, detached from
the otoconiallayer by spontaneous degeneration or
head trauma, gravitate to and settle on the cupula of
the posterior semicircular canal, which is situated
directly inferior to the utricle when the head is
upright. The posterior semicircular canal thus serves
as a receptacle for the detached sediment (Fig. 16.7).
In fact, otoconia are easily dislodged by linear
accelerations or centrifuging in animals (Hasegawa
1933; Igarashi and Nagaba 1968). Otoconial debris
become displaced in old age (Johnson and Hawkins
1972) and lodge either in the posterior semicircular
canals (Vyslonzil 1963) or in the cochlea in
cochleosaccular degeneration (Gussen 1980).
Moriarty et al. (1992) identified basophilic granular
deposits on 22% of the cupulae of 1038 canals of 566
examined individual post-mortem temporal bones.
They determined a higher incidence for the posterior
than for the horizontal or anterior semicircular
canals. Naganuma et al. (1996) found an even high er
percentage of basophilic deposits (horizontal canal:
41 %, posterior canal: 37%, anterior canal: 26%),
which increased with increasing age.
The cupula normally has the same specific gravity
as the endolymph and is a transducer of angular
acceleration only. When heavily loaded, it should
theoretically become sensitive to changes in head
position relative to the gravitational vector (buoyancy

Vertigo

hypothesis, p. 285). The buoyancy mechanism is


used to explain some forms of positional
vertigo/nystagmus which arise after the ingestion of
compounds with differing specific gravities, such as
alcohol (Money et al. 1974), glycerol (Rietz et al.
1987), or"heavy water" (Money and Myles 1974). The
common view was that BPPV simply refiects transformation of the affected cupula from a transducer
of angular acceleration to one of linear acceleration
and (abnormal) angular acceleration, secondary to
the acquired specific gravity differential between the
cupula and endolymph (Gacek 1984; Schuknecht
1969; Rietz et al. 1987). Therefore, in cupulolithiasis
it should be irrelevant on which side of the cupula
the heavy debris become attached (Fig. 16.8).
The traditional view of cupulolithiasis and the
buoyancy mechanism must be incorrect for several
reasons (Table 16.2): BPPV is a positioning rather
than a positional vertigo/nystagmus, because it is
induced only by rapid changes in head position, with
the paroxysmal nystagmus being compatible with
the cupulogram of an ampullofugal stimulation of
the posterior semicircular canal. If cupulolithiasis
were valid, then an ongoing positional vertigol
nystagmus would be expected, as occurs in positional
alcohol nystagmus (Money et al. 1974). The mechanism of cupulolithiasis may account for the subtle
static (persistent) positional nystagmus, which Baloh
and colleagues (1987) observed in 41% of their
patients with BPPV. There was, however, no consiste nt relation between the side of the lesion and the
static positional nystagmus. This lack of directional
correspondence makes interpretation difficult.
Earlier we hypothesised that cupulolithiasis is
mainly causative: the "heavy cupula" causes an oversensitivity of the posterior canal to angular acceleration in the plane specific to the canal (Brandt 1991).
Thus, in a strict sense, BPPV would constitute an

Table 16.2. Cupulolithiasis: pro and contra


Pro
- Single histologieal findings of debris which seemed to be attached to
the cupula
Contra
- No BPPV attacks with slow head tilt (>6 s)
- No typical vertigo with linear head accelerations
- Direction and intensity of induced nystagmus/vertigo do not reflect the
position of"heavy" cu pu la relative to gravity
- Short duration of positioning nystagmus (less than 1 min) with the head
motionless
- Clinical fatiguability with repetitive positioning manoeuvres
- Spontaneous course with varying severity ofthe attacks
- Effieaey of physical therapy with unpredictable remission phases and
relapses
- Incompatible with nystagmus direetion in horizontal BPPV
(From Brandt and Steddin 1993.)

259

Benign paroxysmal positioning vertigo

Fig. 16.8. a Cupulolithiasis: schematic drawing of the cupula in the ampulla of the posterior semicircular canal carrying otoconial
debris.lf this were the pathology causing BPPV, then the static lateral head position brather than the positioning manoeuvre should
be the precipitating factor, irrespective of which side the debris lodge on. Clinically, however, typical BPPV is a positioning vertigo, and
cupulolithiasis of this type cannot explain all clinical features (see p. 253). (From Brandt and Steddin 1993.)

enhanced post-rotatory positioning response, but


not a positional response. The intensity of the positional nystagmus depends on the velo city of the
positioning manoeuvre. BPPV attacks can be avoided
if the challenging position is assumed very slowly
(longer than 6 seconds), wh ich is the case in affected
patients. We no longer believe this explanation, however, because
1. patients would also be sensitive to vertical and

horizontal linear accelerations of the head if it


were correct, and
2. it does not convincingly explain the fatiguability
on repetitive provocation. It is unlikely that with
one or two positioning manoeuvres the detached
partides could disperse from the cupula into the
endolymphatic space and a few hours later again
become attached to the cupula and thereby reactivate the vertigo.

above (p. 253). The diameter of the semicircular canal


(0.32 mm in humans; Curthoys and Oman 1987) is
about one-fifth to one-seventh of the ampulla. A dot
approximately this size in diameter would act as a
plunger on the endolymph and the cupula, if the
slightly turned head is positioned from an upright to
a precipitating lateral position (Fig. 16.9) as soon as
the angle between the canal's plane and the gravity
vector altered. The dot produces pressure or suction
in the canal, thereby deflecting the cupula and eliciting a BPPV attack.
This mechanism would be compatible with

Arguments for canalolithiasis

There is histological proof that inorganic "heavy partides" detached from the otoconiallayer (by degeneration or head trauma) gravitate into the posterior
semicircular canal. Moreover, this is supported by
large series of patients in whom the common dinical
finding indicates that the following conditions figure
in the aetiology of BPPV: head (labyrinthine) trauma,
viral neural labyrinthitis, vertebrobasilar ischaemia,
post-surgery (ear and general), prolonged bedrest due
to unrelated diseases (Gyo 1988) and ageing (Baloh et
al. 1987; Katsarkas and Kirkham 1978).
A dot formed by specific heavy material floating in
the ampullofugal branch of the posterior canal would
be compatible with all five dinical criteria mentioned

a latency of a few seconds (time needed for the


dot-induced flow mechanism to develop by gravitational force),
the ineffectiveness of a very slow positioning
manoeuvre (then the dot would slowly gravitate
along the undermost wall of the canal without
affecting the cupula),
the limited duration of the positioning
vertigo/nystagmus (cupula deflection due to
elastic restoring force ends when the heavy dot
reaches its lowest position in the canal with
respect to the earth surface),
the fatiguability with repetitive provocation
(explained by dispersion of single particles from
the dot, which decreases the plunger effect),
the reactivation of the vertigo after prolonged
bedrest (the result of a new dot being formed by
the particles (Fig. 16.9)),
the direction of the nystagmus during the positioning manoeuvres as explained below.

Attempts were made earlier to attribute BPPV to


floating particles in the semicircular canal (Hall et
al. 1979; Epley 1980a,b; Pagnini et al. 1989; McClure

260

Vertigo

----

Fig.16.9. "Canalolithiasis"; sehematie representation of a free-floating "heavy elot" of otoconial debris aeting like a plunger on
endolymph and eupula of the posterior semicireular eanal, the proposed meehanism for BPPV.ln the normal upright position a the
debris rest at the base of the eupula without any notieeable effeet.lf the head is turned and rapidly positioned to the side in the plane
of the posterior semicireular eanal, the elot, thanks to its greater specifie weight, gravitates downward band, together with endolymph
flow, defleets the eupula in an ampullofugal direetion. When the elot has gravitated to the lowest eurvature of the posterior eanal, vertigo and nystagmus subside beeause the eupula assumes its normal resting position c.lf the patient returns to the seated position, a
similar effeet could explain the reversed direetion of nystagmus and vertigo. (For explanation of fatiguability due to repeated provoeation and physieal therapy, p. 259.) (From Brandt and Steddin 1993.)

1988). We believe there are now convincing arguments that canalolithiasis and a clot-induced
endolymph flow mechanism are the significant
causative factors. Despite ongoing discussions about
possible vestibular habituation effects (Steenerson
and Cronin 1996; Smouha 1997), canalolithiasis is
the only mechanism that explains the success of the
highly effective physical therapy by positioning
manoeuvres, as proposed by Brandt and Daroff
(1980), Semont and colleagues (1988), Epley (1992),
or Lempert et al. (1996), who demonstrated the efficacy of canal-clearing manoeuvres by performing
backward rotation of the posterior canal during the
use of a flight simulator.
As Fig. 16.10 shows, we believe that the floating dot
of particles can be sluiced down by both of the
described positioning manoeuvres via the upper
ampullofugal branch of the posterior canal into other
labyrinthine recesses where they are no longer
causative. The mechanism demonstrated in Fig. 16.10
coincides with the hitherto unexplained observation
(Pace-Balzan and Rutka 1991; Semont et al. 1988;
Husler and Pampurik 1989) that after a positioning
manoeuvre from the position in Fig. 16.10(b) (with
the affected ear undermost) to that in Fig. 16.10(c)
(with the affected ear uppermost), the induced nystagmus still rotates toward the uppermost ear.
Cupulolithiasis predicts an ampullopetal deflection of
the cupula, which results in nystagmus toward the
undermost ear. The unexpected direction, however,
can be easily explained by the dot-induced
endolymph flow mechanism, which acts in the same
direction in the two different positioning manoeuvres
in Fig. 16.10 (b and c). We interpret the direction of
the nystagmus thus induced by the positioning manoeuvre as indirect proof that canalolithiasis is valid
(Table 16.3).

Table 16.3. Canalolithiasis: pro and contra


Pro
- Compatible with all dinical features of BPPV-induced nystagmus and
with all arguments against cupulolithiasis (see Table 16.2)
- The positioning nystagmus toward the uppermost ear as induced by
1800 head tilt (Fig. 16.10) provides indirect proof of canalolithiasis
- Compatible with features of horizontal BPPV nystagmus
-Intraoperative findings of free-floating endolymphatic partides in
symptomatic patients
Contra
- Histological findings of deposits in the semicircular canals in
asymptomatic patients
- Sudden onset of the disease (no slow build-up of dot and symptoms?)
(From Brandt and Steddin 1993.)

Free-floating endolymphatic particles were first


found intraoperatively during posterior semicircular
canal occlusion by Parnes and McClure (1992). In a
prospective study on 26 patients undergoing the
posterior canal occlusion procedure and 76 patients
undergoing labyrinthine surgery for vestibular
schwannoma or labyrinthectomy, particulate matter
was observed in eight of 26 patients with BPPV; no
particles were observed in any of the 76 patients
(Welling et al. 1997).
The possibility of a combination of canalolithiasis
with cupulolithiasis has been demonstrated for hBPPV (Steddin and Brandt 1996; p. 274). There are
positions of the head in which the free-floating clot
should settle on the cupula, and subsequently after
initial canalolithiasis, cupulolithiasis should occur,
according to the bouyancy mechanism. Do some of
the patients with intractable BPPV have precipitate
settled on the cupula? If this is the case, the nystagmus pattern should be different from that in treatable cases.

261

Benign paroxysmal positioning vertigo

Fig.16.10. Schematic drawing ofthe dot of otoconial debris dispersed and sluiced by positioning manoeuvres from the normal a to
the challenging b position and to the opposite side c. Depending on the change of the head position relative to the gravitational vector, the dot settles to the lowermost part of the canal, and after transition from position b to c, leaves the canal in order to enter other
labyrinthine recesses where it no longer causes vertigo attacks. This scheme demonstrates why physical therapy with positioning
manoeuvres is effective when the dot or the debris leave the affected semicircular canal. The direction of induced nystagmus in c indirectly proves that canalolithiasis rather than cupulolithiasis is the significant mechanism. Only with a free-floating dot is the direction
of positioning nystagmus in c the same as in b. (From Brandt and Steddin 1993.)

Suzuki et al. (1996) described a functional animal


model of BPPV using an isolated frog semicircular
canal. When otoconia were place on the cupular surface to mimic cupulolithiasis ampullary nerve action
potential instantaneously responded; when otoconia
were dropped into the canal to mimic canalolithiasis
action potentials responded with the otoconial flow
after a latent period.
Gordon (1992) proposed an alternative air-bubble
theory. As appealing and simple as his theory is, it is
not supported by all dinical characteristics of BPPV,
as is canalolithiasis (Brandt and Steddin 1992).
Movement of bubbles inside the membranous
labyrinth is compatible with some features but not
with others, such as positioning nystagmus toward
the uppermost ear when the patient performs a 180
body tilt (Fig. 16.11).
Unilateral mimicking bilateral BPPV

On the basis of our observations and the assumption


of a free-floating heavy dot (canalolithiasis), headpositioning manoeuvres were described in which
unilateral BPPV mimics bilateral BPPV (Steddin and
Brandt 1994b).
The proposed mechanism of canalolithiasis not
only explains the well-known features of BPPV but
also predicts the direction and duration of rotational
vertigo and nystagmus elicited by various head positions or positioning manoeuvres. For example, what
happens if a patient sitting in anormal upright position but with his or her head turned 45 toward the

unaffected ear is tilted toward the side of the unaffected ear (Fig. 16.12)? The hypothesis predicts that
the free-moving dot will come to rest on the cupula
of the posterior semicircular canal, thereby deflecting the cupula in the ampullopetal direction, as in
cupulolithiasis. This should cause an inhibition of
both the superior oblique eye musde on the affected
side and the inferior rectus eye musde on the unaffected side, resulting in a geotropic rotatory nystagmus to the side of the unaffected ear, which is
undermost in this position. These considerations are
necessary for interpretation of diagnostic test
manoeuvres.
Figure 16.13 shows what should happen in a
patient with left-sided typical BPPV based on the
mechanism proposed above. The heavy dot (Fig.
16.13, top) is located at the ampullofugal side of the
ampulla of the left posterior semicircular canal. This
causes no stimulation in the normal upright position. If the patient is quickly tilted toward the affected
left ear without concurrent head rotation (nose in
horizontal plane after tilt), the dot gravitates toward
the lower part of the canal, resulting in canalolithiasis with a typical BPPV attack (Fig. 16.13, middle).
However, if the patient is tilted with the same alignment of the head and neck with the trunk toward the
opposite, right side, the posterior canal of the affected
ear - which is now uppermost - mayaiso become
stimulated by the dot. In this head position, the
plane of the posterior canal is tilted vertically by
approximately 45. The dot gravitates to and
becomes settled on the cupula, but now (because of

262

Vertigo

Fig.16.11. Schematic representation of the two theories in question (canalolithiasis versus air bubbles) to explain the pathomechanism
of BPPV. Position of the head and body is depicted (top) as weil as the corresponding spatial orientation of the posterior semicircular canal
(middle) and the assumed movement of a heavy dot in canalolithiasis (black) or of endolymphatic air bubbles (AB), causing typical pos itioning nystagmus (bottom).Arrows indicate affected ear, movement of dot or bubbles, subsequent cupula deflection, and nystagmus
direction. Canalolithiasis.A free-floating heavy dot acts like a plunger on endolymph and cupula of the posterior semicircular canal (middie), the proposed mechanism for BPPV. The dot, because of its greater specific weight, gravitates downward and, by endolymph flow,
deflects the cupula in an ampullofugal direction.lf the head is tilted by 180 0 to the opposite side (right), according to the changes of the
head position relative to gravitation, the dot settles to the lowermost part and leaves the canal in order to enter other labyrinthine recesses (explanation for the efficacy of the deli berate manoeuvres).The direction of induced nystagmus indirectly proves that canalolithiasis is
a significant mechanism. Only in the ca se of a free-floating dot is the direction of positioning nystagmus toward the right ear in both
manoeuvres. Air bubble theory. Free-floating air bubbles are compatible with the BPPV nystagmus with the head tilted toward the affected ear (middle) but not with the head tilted 180 0 toward the unaffected ear (right). The movements of air bubbles and the heavy dot are
then opposite in direction and accordingly would cause different cupula deflections. (From Brandt and Steddin 1992.)

Fig.16.12. Schematic drawing of"transient cupulolithiasis" caused bya free-floating "heavy dot" in the ampullofugal branch of the
left posterior semicircular canal. Diagram shows the effect of body tilt from the upright position (top) to the right side (bottom) in a
patient with left-sided p-BPPV, with the left posterior canal parallel to the plane of body tilt. With the left ear uppermost, the free-floating dot deflects the cupula ampullopetally toward the utride, causing an inhibitory stimulation (minus sign) of the ipsilateral superior
oblique eye musde and the contralateral inferior rectus eye muscle.Arrows indicate fast phase of nystagmus. (From Steddin and
Brandt 1994b.)

Benign paroxysmal positioning vertigo

gravity) it defiects the cupula ampullopetally (Fig.


16.l3, bottom) . This may result in a positional vertigo
and nystagmus toward the undermost (unaffected)
ear. Theoretically, particles of different densities
gravitating in the canal should be more effective
than an equal weight impinging directly on the
cupula, because of the cross-sectional differential
which is approximately 1:5 to l.7 for the diameters of
the semicircular canal and the ampulla (Curthoys
and Oman 1987; Epley 1995).
Hence, the characteristics of this positional nystagmus (cupulolithiasis mechanism) are different
from those of the positioning nystagmus on left head
tilt (canalolithiasis) in that the nystagmus is not
paroxysmal but on-going with lower amplitude and
at low frequency (patients report mild rotational or
indefinite vertigo).
Since the nystagmus has the same direction as in
true BPPV of the unaffected ear, it could be misinterpreted as an asymmetrical bilateral BPPV, which is
more pronounced for the left than for the right ear.
An inadequate therapeutic approach would result.
The above-described mechanism may account in
part for the persistent positional nystagmus
observed by Baloh et al. (1987) in 41 % of their
patients. Furthermore, Dingle et al. (1992) were
puzzled by a patient with suspected bilateral BPPV
whose symptoms immediately and completely

263

resolved following unilateral canal occlusion of the


posterior semicircular canal on the "worse affected"
side. Parnes and McClure (1992) also reported that a
patient with bilateral BPPV remained free of BPPV
after canal occlusion on the "more symptomatic
side". Our own experiences confirm those findings:
patients with seemingly bilateral BPPV were relieved
simultaneously of all symptoms after performing
liberatory manoeuvres designed for unilateral BPPV.
Beyond the classic Dix and Hallpike manoeuvre,
we recommend a sequence of positioning manoeuvres for the definitive diagnosis of true bilateral pBPPV. The spatial orientation of the left and right
posterior semicircular canals du ring and after lateral
head-trunk tilt is fundamental. The plane of the suspected posterior semicircular canal must be adjusted
coplanar to the plane in which the l-ateral tilt is performed: head rotated by 45 to the left shoulder with
tilt to the right or head rota ted by 45 to the right
shoulder with body tilt to the left, respectively. This
guarantees maximum excitation of the undermost
posterior semicircular canal by ampullofugal defiection of the cupula, whereas the contralateral posterior
semicircular canal - with its plane nearly perpendicular to the plane of body tilt - is not stimulated.
Figure 16.14 shows the respective manoeuvres in
a patient with suspected unilateralleft-sided BPPV.
While the patient is in a sitting position, his/her

Fig.16.13. Unilateral mimicking bilateralleft-sided p-BPPV.lf the frontal plane of the patient's head is parallel to the plane of body
tilt, a free-floating clot may cause geotropic nystagmus in either the ipsilateral (middle) or contra lateral (bottom) position by
ampullofugal (middle; canalolithiasis mechanism) or ampullopetal (bottom; cupulolithiasis mechanism) deflection of the left posterior
semicircular canal's cupula.Arrows indicate fast phase of nystagmus; plus sign, excitatory stimulus and minus sign inhibitory stimulus.
(From Steddin and Brandt 1994b.)

Vertigo

264

pendicular to the plane of body tilt throughout the


entire manoeuvre; the clot remains in its former
position without affecting the cupula. If a BPPV
attack were to occur, it would have to be elicited by
the undermost right posterior semicircular canal
(open circle represents the floating clot), which
proves bilateral BPPV.

Aetiology

Fig.16.14. (Top) diagnostic test for left-sided p-BPPV. With the


head turned 45 toward the right shoulder, an attack is elicited by
a dot within the left posterior canal (small solid circle).A dot within the right posterior canal (sm all open circle) has no effect,
because the canal's plane is perpendicular to the plane of body
tilt. (Bottom) diagnostic test for right-sided p-BPPV. With the head
turned 45 toward the left shoulder, an attack is elicited by a dat
within the right posteriar canal (small open circle).A dat within
the left posterior canal (sm all solid circle) has no effect. (Fram
Steddin and Brandt 1994b.)

head is turned 45 about the z-axis to the right (Fig.


16.14, top Zeft) before he or she is tilted to the left
(Fig. 16.14, top right). A BPPV attack with geotropic
rotational nystagmus is direct proof of a left-sided
BPPV, because vertigo cannot be elicited in this position by the right posterior semicircular canal (with
its plane perpendicular to the plane of body tilt).
After the patient has sat up and remained in this
position for some time, the patient's head is turned
45 about the z-axis to the left (Fig. 16.14, top Zeft)
and then tilted quickly to the right (Fig. 16.14, top
right). A unilateralleft-sided BPPV (soZid circle represents the floating clot) cannot cause vertigo or
positional nystagmus in this position, because the
plane of the left posterior semicircular canal is per-

Particles have been frequently found in the membranous labyrinth of symptomatic and asymptomatic patients. These particles seem to be identical
to otoconia or otoconial (calcite) fragments (Gussen
1980; Kveton and Kashgarian 1994). The particulate
matter from within the membranous posterior semicircular canal from a patient at the time of canal
occlusion for intractable BPPV was examined by
scanning electron microscopy. It appeared to be
morphologically consistent with degenerated otoconia
(Welling et al. 1997). However, there is still some
uncertainty about the origin of these deposits, and it
seems likely that more than one possible explanation
is needed to account for their existence (Moriarty et
al. 1992).
Large series of patients (Katsarkas and Kirkham
1978; Baloh et al. 1987) provided support for the
common clinical finding that the following playa
role in the aetiology of BPPV: head (labyrinthine)
trauma, vestibular neuritis, vertebrobasilar
ischaemia, postsurgery (ear and general), prolonged
bedrest due to unrelated diseases, and most often
"idiopathic" conditions (e.g. ageing). Single case
descriptions include postoperative bedrest (Gyo
1988) or neurosurgical removal of an osteoma using
hammer and chisei (Andaz et al. 1993). In the early
stages, BPPV is usually experienced on awaking in
the morning rather than on first lying down. In the
series of 240 patients described by Baloh et al.
(1987), the origin was idiopathic in about half of the
cases. In the remainder the most commonly identified causes were head trauma (17%) and vestibular
neuritis (15%). When patients present with posttraumatic BPPV (Gordon 1954; Barber 1964), it is
sometimes difficult to determine retrospectively
whether the trauma caused the vertigo or vice versa.
In less than 10% of patients, BPPV is bilateral (mostly
asymmetrieal); this is particularly more frequent in
post -traumatic cases (Longridge and Barber 1978).
We found that 12% of a total of 104 patients with
unilateral BPPV had suffered from vestibular neuritis
days or years previously (Bchele and Brandt 1988).
The relatively frequent occurrence of BPPV following vestibular neuritis was first attributed to
ischaemia of the anterior vestibular artery (Lindsay

Benign paroxysmal positioning vertigo

265

and Hemenway 1956), but it is more likely to be due


to viral inammation at the site of the vestibular
nerve (see Vestibular neuritis, p. 67). Vestibular neuritis is a partial rather than complete unilateral
vestibular paresis (Bchele and Brandt 1988), since
the resulting BPPV requires that the function of the
posterior canal be preserved.
In a retrospective population-based study in
Minnesota, the age- and sex-adjusted incidence of
BPPV was 64 per 100000 population per year
(Froehling et al. 1991). Incidence increased by 38%
with each decade of life. From epidemiological surveys, the incidence of BPPV in Japan was estimated
to be between ll-17 per 100000 population
(Mizukoshi et al. 1988). In a retrospective study of
806 patients 70 years of age or older who complained
of dizziness, 41 % had a history strongly suggestive of
BPPV (Bloom and Katsarkas 1989). The age of onset
of BPPV ranges from adolescence to old age, and in
the idiopathic group exhibits a peak incidence in the
sixth and seventh decades, but onset tends to be earlier on the average in symptomatic forms of BPPV. In
the idiopathic group females exceed males by 2 to 1
(Katsarkas and Kirkham 1978; Baloh et al. 1987),
whereas the sexes are about equally distributed in the
post-traumatic and post-vestibular neuritis forms.

Positional exercises and liberatory manoeuvres

Fig.16.15. Positional exereises for effeetive physieal therapy


for BPPV as proposed by Brandt and Daroff in 1980. Patients are
instrueted to sit and then to move rapidly into the ehallenging
position, to remain in the position for at least 30 s, and then to sit
up for 30 s before assuming the opposite head-down position for
30 s. These exercises are repeated serially 5-10 times a day.

The positional exercises proposed in 1980 (Brandt


and Daroff 1980) were the first effective physical
therapy (Fig. 16.15). The exercises were a sequence
of rapid lateral head/trunk tilts, repeated serially to
promote dispersion of the debris toward the utricular cavity. We instructed the patients to sit; to then
move rapidly into the challenging position to induce
the correct plane-specific stimulation of the posterior semicircular canal; to remain in the position until
the evoked vertigo subsided, or for at least 30 s; and
then to sit up for 30 s before assuming the opposite
head-down position for an additional 30 s. Troost
and Patton (1992) reviewed and diagrammed this
exercise protocol.
The Semont and Epley liberatory man oeuvres
require only a single sequence, making them preferable to the multiple repetitions over many days
required by the Brandt-Daroff exercises. With
canalolithiasis as the established mechanism of
BPPV, we can now explain the efficacy of the therapies
according to anatomic and physical principles.
Figure 16.16 illustrates the Semont manoeuvre in a
patient with typical (posterior canal) left-sided BPPV.

The elot causes no deection of the cupula in the


upright position. When the patient is quickly tilted
toward the affected left ear with a 45 head rotation to
the right (moving the left posterior canal to a plane
corresponding to the plane of the head tilt), the elot
gravitates toward the lower part of the canal, causing
the cupula to deect downward (ampullofugal), and
triggering a typical BPPV attack. These events explain
the latency of a few seconds (the time needed for the
elot-induced endolymph ow to develop by gravitational force ), the ineffectiveness of a very slow positioning movement (the elot would then slowly
gravitate along the undermost wall of the canal without plugging the canal and deecting the cupula), and
the short duration of the positional vertigo/nystagmus (the cupula deection ends when the elot reaches
its lowest position in the canal) (Brandt and Steddin
1993). If the patient is swung toward the opposite
right side with the nose down, the elot will gravitate
downward, causing stimulation of the posterior canal
of the affected left ear (now uppermost). If no vertigo
and nystagmus are elicited, we gently shake the

Management

266

Vertigo

RE

LE

LE

RE

RE

LE

Fig.16.16. Schematic drawing of the Semont liberatory manoeuvre in a patient with typical BPPV of the left ear. Boxes from left to
right: position of body and head, position of labyrinth in space, position and movement of the clot in the posteriar canal and resulting
cupula deflection,and direction of the rotatory nystagmus. The clot is depicted as an open circle within the canal; a black circle represents the final resting position of the clot. (1) In the sitting position, the head is turned horizontally 45 to the unaffected ear. The clot,
which is heavier than endolymph, settles at the base of the left posterior semicircular canal. (2) The patient is tilted approximately 105
toward the left (affected) ear. The change in head position, relative to gravity, causes the clot to gravitate to the lowermost part of the
canal and the cupula to deflect downward, inducing BPPV with rotatory nystagmus beating toward the undermost ear.The patient
maintains this position for 3 min. (3) The patient is turned approximately 195 with the nose down, causing the clot to move toward
the exit of the canal. The endolymphatic flow again deflects the cupula such that the nystagmus beats toward the left ear, now uppermost. The patient remains in this position far 3 min. (4) The patient is slowly moved to the sitting position; this causes the clot to enter
the utricular cavity. Abbreviations: A, P, and H = anterior, posterior, horizontal semicircular canals. Cup = cupula, UT = utricular cavity, RE
= right eye, and LE = left eye. (From Brandt et al. 1994.)

patient's head in this position; this sometimes seems


to facilitate loosening and gravitation of the dot. The
patient is then slowly moved to the upright position;
the dot will gravitate downward through the common
crus of the posterior and anterior canals and enter the
utricular cavity, where it becomes harmless. We share
the experience of others (Serafini et al. 1996) that
complete recovery after a single manoeuvre is
achieved in about 50-70% of cases. Semont et al.

(1988) recommended having the patient maintain the


upright position for 48 ho urs following the liberation,
but we have not found this to be necessary.
Figure 16.17 illustrates the Epley manoeuvre
(1992) as modified by Herdman et al. (1993) and
others (Harvey et al. 1994) in a patient with typical
(posterior canal) left-sided BPPV. The dot causes no
deflection of the cupula in the upright position with
the head turned horizontally 45 to the affected ear.

267

Benign paroxysmal positioning vertigo

~.
RE

LE

RE

LE

RE

LE

RE

LE

Fig.16.17. Schematic drawing of modified Epley liberatory manoeuvre. Patient characteristics and abbreviations are as in Fig. 16.16.
(1) In the sitting position, the head is turned horizontally 45 to the affected (Ieft) ear. (2) The patient is tilted approximately 105 backward into a slight head-hanging position, causing the dot to move in the canal, deflecting the cupula downward, and inducing the
BPPV attack. The patient remains in this position for 3 minutes. (3a) The head is turned 90 to the unaffected ear, now undermost, and
(3b) the head and trunk continue turning another 90 to the right, causing the dot to move toward the exit of the canal. The patient
remains in this position for 3 minutes. The positioning nystagmus beating toward the affected (uppermost) ear in positions 3a and 3b
indicates effective therapy. (4) The patient is moved to the sitting position. (From Brandt et al. 1994.)

When the patient is quickly tilted backward into a


slight head-hanging position, the dot gravitates
downward in the posterior canal, deftecting the
cupula downward and inducing a BPPV attack.
Rotation of the head and trunk toward the unaffected

right ear causes further movement of the dot downward (ampullofugal) toward the exit of the canal,
resulting in positioning vertigo and nystagmus
toward the affected (now uppermost) ear. The final
uprighting of the patient causes the dot to enter the

Vertigo

268

utricular cavity, where it becomes harmless. Li


(1995) found that the success rate improved if the
procedure is combined with mastoid vibration; this
corresponds to our attempts to promote "canalclearing" by additional head shaking (p. 265). In our
opinion, the frequently used term "canalith repositioning manoeuvre" (Epley 1992) is incorrect, since
it is unlikely that the clot is "repositioned" to its originallocation. After an effective liberatory manoeuvre,
however, some patients complain of a transient
"otolithic vertigo" with unsteadiness lasting for days.
This could be explained by the unequal heavy loads
of both otoliths when the clot settles on the macula
bed.
Following effective physical liberation, approximately 50% of patients (Baloh et al. 1987) will
experience a recurrence of attacks; 10% to 20%
occur in the first 2 weeks (Herdman et al. 1993). The
recurrences may be due to re-entry of the debris into
the posterior canal from the utricular cavity and
should be treated with the same manoeuvre that
induced resolution of the initial episode.
The process illustrated in Figs. 16.16 and 16.17
explains the seemingly paradoxical observation
(Semont et al. 1988; Pace-Balzan and Rutka 1991;
Husler and Pampurik 1989) that the finalliberatory
positioning with the affected ear uppermost (Fig.
16.16, panel 3; Fig.16.17, pane13b) induces nystagmus
that beats toward that ear (Brandt and Steddin 1993).
As described above, cupulolithiasis predicts an
ampullopetal deflection of the cupula that would
cause nystagmus to beat toward the undermost ear,
whereas in canalolithiasis, the clot-induced
endolymphatic flow causes ampullofugal deflection
of the cupula and nystagmus beating to the uppermost ear. Moreover, the upward direction of the nystagmus induced by the final positionings is a
clinically relevant observation in that it provides
reasonable certainty that the clot has exited the canal
(or will so exit in the modified Epley manoeuvre)
and the patient will be free of symptoms ("liberated").
If the nystagmus does not beat upward toward the
affected ear, the clot is probably still inside the canal;

if the nystagmus beats downward toward the unaffected ear, the clot must have moved toward the
cupula, causing an ampullopetal deflection (Fig.
16.18). In either situation, the procedure should be
repeated. If the nystagmus fails to beat upward following the second procedure and the BPPV persists,
we schedule areturn visit for the same manoeuvre. If
the second session fails, we try a different liberatory
manoeuvre (i.e. modified Epley, if we first used
Semont, or vice versa). If both liberatory manoeuvres fail, we prescribe Brandt -Daroff exercises.
A possible complication of liberatory manoeuvres
is that the clot leaves the posterior canal but instead
of staying in the utricular cavity enters the anterior
(via common crus) or the horizontal canal. Thus,
p-BPPV may convert to h- or a-BPPV. This occurred
in 5 of 85 patients originally with typical p- BPPV
(horizontal canal: 3, anterior canal: 2) after they had
undergone liberatory manoeuvres (Herdman and
Tusa 1996). "Canalith jam" is another speculative
description of hitherto unexplained transient
phenomena that rarely occur during physical treatment (Epley 1995): ''An interesting phenomenon that
I have occasionally observed while undertaking the
canalith repositioning procedure is a sudden conversion of transient nystagmus to a rapid form that persists irrespective of head position. Simultaneously
the patient usually complains of intense vertigo. I
believe the mechanism to be a jamming of the
canaliths when migrating from a wider to a narrower segment (i.e. from ampulla to canal or at the
bifurcation of the common crus). For treatment the
crus is repositioned (inverted) and vibration is
applied. Gravity backs dense debris out of the jam."
In the rare case of anterior canal BPPV, the spontaneous symptoms also occur when the affected ear is
uppermost. The liberatory manoeuvres seem to work
here (Baloh et al. 1993), but, at least theoretically, they
should also be initiated with the abnormal ear uppermost. In the less rare case of horizontal canal BPPV,
Baloh et al. (1993) found that Brandt-Daroff exercises
may be more effective than the liberatory manoeuvres.
The Semont and the modified Epley man oeuvres

LE

RE

Fig.16.18.

Schematic drawing of an ineffective Semont liberatory manoeuvre to be compared with Figure 16.16, panel 3. After the
patient is tilted to the right, the dot migrates back toward the cupula. Endolymph flow causes an ampullopetal cupula deflection with the
nystagmus beating downward toward the unaffected ear.This indicates that the liberatory manoeuvre has probably failed. (From Brandt et
al. 1994.)

269

Benign paroxysmal positioning vertigo

have a theoretical advantage over each other. Epley's


final head positioning (Fig. 16.17, panel 4) causes the
clot to migrate further from the cupula and closer to
the canal exit, whereas Semont's movement (Fig.
16.16, panel 3) reaches a higher acceleration and
thus is more likely to move the clot. These relative
advantages may cancel out, since similar efficacy is
achieved by both manoeuvres (Herdman et al. 1993).
In conclusion, there are three highly effective
therapies for the common posterior canal BPPV. The
Semont and modified Epley liberatory manoeuvers
are easily performed by a physician or trained
therapist, and the instructions for horne use of the
Brandt-Daroff exercises are simple. With both liberatory manoeuvres, nystagmus beating toward the
uppermost affected ear confidently predicts
therapeutic success.

Surgical procedures
In those rare patients who do not respond even to
appropriate and prolonged physical therapy

surgical plugging of the posterior semicircular


canal via a transmastoid approach or
surgical transection of the posterior ampullary
nerve via amiddie ear approach can be considered.

In our experience with more than 1000 patients with


typical BPPV, however, only a few individual patients
did not respond to physical therapy, and they ultimately required selective surgical transection or canal
plugging. We believe that an indication for surgical
intervention is still too frequently complied with
before the possibilities of physical therapy are completely exhausted. This view is shared by Epley
(1995), who has invented his own effective liberatory
procedure. He believes that the disability ensuing
from multiple, unpredictable recurrences is over the
long term a more common indication for surgery.
Singular neurectomy

Transection of the posterior ampullary nerve provides relief of vertigo; it is, however, not easy to locate
surgically the relevant semicircular canal (Leuwer and
Westhofen 1996), and sensorineural hearing loss is a
possible complication among several others (Gacek
1978, 1984; Epley 1980a,b). In his series of 137
patients, Gacek (1995) found complete relief of vertigo in 94% with sensorineural hearing loss in 3%.
Plugging of the posterior semicircular canal

Parnes and McClure (1990,1991) first described


transmastoid posterior semicircular canal occlusion

with complete relief of BPPV and preserved lateral


semicircular canal function as a simpler and safer
alternative to singular neurectomy. Others have confirmed the success of fenestration and occlusion of
the posterior canal (Pace-Balzan and Rutka 1991;
Hawthorne and El-Naggar 1994; Kanayama et al.
1995). Possible complications are surgicallabyrinthitis with reversible ataxia and sensorineural hearing
loss, which also mostly recovers, or inadvertent plugging of neighbouring (e.g. horizontal) semicircular
canals. Arai et al. (1989) observed direction-changing positional nystagmus that continued for several
months after vertical canal plugging in monkeys.
Modifications of the procedure using laser techniques have also been reported (Anthony 1991, 1993;
Kartusch and Sargent 1995; Nomura et al. 1995), but
convincing proof of their superiority is lacking. It
should be noted here that in 1950 Vogel was the first
not only to speculate that endolymph-flow could be
the cause of BPPV attacks, but also to propose
the plugging of semicircular canals as a possible
treatment.
Table 16.4 summarises information given in this
chapter about p-BPPV.

Horizontal semicircular canal BPPV


(h-BPPV)
The first cases ofh-BPPV were described by McClure
(1985). Later, Pagnini et al. (1989) and Baloh et al.
(1993) presented more detailed clinical descriptions
of this condition. Now h-BPPV accounts for about
10-20% of all patients presenting with BPPV. It may
be combined with p-BPPV of the same or the contralateral ear. Transitions between h-BPPV and
p-BPPV are also possible, particularly as a result of
therapeutic positioning manoeuvres. Whereas typical h-BPPV is caused by canalolithiasis, atypical
h-BPPV may occur with apogeotropic positioning
nystagmus caused by cupulolithiasis. The aetiological
factors are the same as in p-BPPV.
Patients do not report experiencing episodic vertigo when getting in or out of bed, but when rolling
the head from side to side while supine.
Consequently, positioning testing of BPPV patients
should include the sitting-to-lateral head positioning manoeuvre for p-BPPV and the supine-to-lateral
head rotation manoeuvre for h-BPPV (Fig. 16.19).
The most effective physical therapy seems to be a
forced prolonged bedrest, during which the affected
ear remains uppermost. Relapses ofh-BPPV seem to
occur more frequently than relapses of p-BPPV.

Vertigo

270
Table 16.4. Benign paroxysmal positioning vertigo (typical posterior
semicircular canal type, p-BPPV)

Clinical syndrome
- Brief attacks of rotational vertigo and concomitant rotatory-linear
nystagmus precipitated by rapid head-trunk tilt toward the affected ear
or by neck extension (when first Iying down in bed, sitting up from a
supine position, turning over in bed from one side to the other,
extending the neck to look up)
- Latency:
Vertigo and nystagmus begin 1 or more seconds
after head tilt
- Duration:
Attacks last less than 40 s
- Nystagmus:
Linear-rotatory, with the fast phase beating
toward the undermost ear or upward
- Reversal:
When the patient returns to the seated position,
vertigo and nystagmus reoccur in the opposite
direction
- Fatiguability:
Repetition of the manoeuvre results in everlessening symptoms
Incidence/age/sex
Most common cause of vestibular vertigo that manifests throughout
life preferably in the elderly.lncidence 11-64 per 100000 population
per year with a peaking incidence in the sixth and seventh decades;
females exceed males by 2 to 1.
Pathomechanism
"Canalolithiasis" of the posterior semicircular canal; dislodged
otoconia (degeneration, trauma) congeal to form a free-floating
"heavy" do!, which always gravitates to the most dependent part of
the canal during changes in head position, thereby causing push or
pull forces on the cu pu la

ofh-BPPV is based on the following features (Strupp


et al. 1995):

Aetiology
- "Idiopathic" forms (ageing) - 50%
- Symptomatic forms - 50% (e.g. head trauma, vestibular neuritis,
prolonged bedrest)
Course/prognosis
- Natural history is considered benign because it resolves spontaneously
within days to months in most patients, persists in about 20-30% when
untreated and recurs in 30-50% after variable periods for years.
Management
- Physicalliberatory manoeuvres to free the canal of the "heavy" dot
- Brandt and Daroff manoeuvres
- Semont manoeuvre
- Epley manoeuvre
are successful in almost 100% of the patients within days or weeks.
- Surgical procedures (plugging of the posterior canal or posterior
ampullary nerve section) are effective but unnecessary in all but a few
exceptional cases.
Differential diagnosis
Central positional vertigo/nystagmus, vestibular paroxysmia,
perilymph fistula, drug or alcohol intoxication, Meniere's disease,
psychogenic vertigo

The clinical syndrome


The clinical characteristics of the horizontal semicircular canal variant of BPPV are easy to distinguish
from those of posterior canal BPPV (McClure 1985;
Pagnini et al. 1989; Baloh et al. 1993). The diagnosis

The patient has a history of brief episodes of vertigo, usually induced by rolling the head from
side to side while supine.
Positioning testing reveals a linear horizontal
nystagmus toward the undermost ear (geotropic)
when the head of the supine patient is rapidly
turned from side to side around the longitudinal
z-axis (barrel roll).
Horizontal positioning nystagmus with the head
turned to either side beats stronger toward the
affected ear (Fig. 16.20); when this position is
maintained, nystagmus often reverses its direction (Figs. 16.21 and 16.22).
Positioning nystagmus in h-BPPV exhibits short
latencies 5 s) and lasts longer (20-60 s) than in
p-BPPV.
h-BPPV rarely fatigues with repetitive positioning manoeuvres.
About one-third of the patients show moderate,
horizontal semicircular paresis du ring caloric
irrigation of the affected ear.
Positioning vertigo attacks are often more severe
than in p-BPPV and more frequently associated
with nausea. As distinct from p-BPPV, attacks are
not elicited by the patient getting in or out of
bed, bending over, or extending the neck.
However, some patients report brief episodes of
vertigo when turning their head while erect
(Baloh et al. 1993).

Atypical h-BPPV with apogeotropic position al


nystagmus

Atypical cases of h-BPPV have been described in


which the positional rather than positioning nystagmus beats toward the uppermost ear (Agus et al.
1995; Baloh et al. 1995; Nuti et al. 1996). They have
been identified as horizontal canal "cupulolithiasis"
as distinct from typical "canalolithiasis" (Baloh et
al. 1995; Steddin and Brandt 1996). In these exceptional patients episodic vertigo and nystagmus can
be termed positional, since they depend on the
head position relative to gravity rather than on the
positioning manoeuvre. Vertigo and nystagmus
may be more or less severe than in canalolithiasis.
The following are the typical differential diagnostic
criteria (Baloh et al. 1995; Steddin and Brandt
1996):

The direction of positional nystagmus is toward


the uppermost ear (apogeotropic).
Nystagmus may last for minutes when the precipitating position is maintained.

Benign paroxysmal positioning vertigo

271

Fig.16.19. Precipitating positioning manoeuvres for horizontal semicircular canal BPPV: the head of the supine patient is rapidly
turned from side to side around the longitudinal z-axis (barrel roll).

Nystagmus depends only on the assumed head


position rather than the net angle of head rotation.
Thus, positional nystagmus beating toward the
uppermost ear is not a pathognomonic sign of central vestibular disturbance (p. 292). However, it can
indicate occasional cupulolithiasis of the horizontal
semicircular canal.

Natural course
h-BPPV is reported to occur from the third to the ninth
decade, with a mean age of manifestation in the sixth
decade (Nuti et al. 1996; De la Meilleure et al. 1996).
There is a slight preponderance of cases in females. The
right ear is affected as frequently as the left. The onset
of the condition is usually abrupt as is its spontaneous
remission after days to weeks. There are rare persistent
courses which last for months to years. We share the
experience of Baloh et al. (1993) and De la Meilleure et
al. (1996): about half of the patients have a few or multiple relapses of the same condition. This is also true far
the h-BPPV variant of cupulolithiasis.

Aetiology and pathomechanism


The aetiology of h-BPPV is the same as that of pBPPV (Baloh et al. 1993; De la Meilleure et al. 1996).
It is "idiopathic" in most cases, post-traumatic in
about 20%, or the result of physical liberatory

manoeuvres for treating p-BPPV (transition from


p-BPPV to h-BPPV), general or temporal bone
surgery, or prolonged bedrest.
In an attempt to explain the clinical features of
h-BPPV, McClure (1985) proposed canalolithiasis as
the underlying mechanism, a theory that was backed
by Baloh et al. (1993). This concept is strongly supported by the intensity of the positioning nystagmus: it is maximal when a patient with h-BPPV of
the right horizontal semicircular canal turns his
head about the longitudinal z-axis from the left lateral to the right lateral position while supine, and it
is much slower when he turns his head to the right
lateral position while in a supine position (nose up)
(Fig. 16.20; Strupp et al. 1995). The fact that this difference in intensity of nystagmus depends on the
initial head position and direction of head rotation
indirectly proves that the clot moves freely (to and
fro) within the segment of the horizontal semicircular canal diametrically opposite the ampulla
(Fig. 16.23). Canalolithiasis is compatible with all
clinical features of typical h-BPPV.
Much higher slow-phase velocity nystagmus
occurs when the head is rotated to the affected side.
We measured a particularly high slow-phase velo city
(175-370 0 /s) for ampullopetal displacement (Figs.
16.20 and 16.24). Ewald's second law may explain the
asymmetry between ampullopetal and ampullofugal
displacement (Halmagyi et al. 1990; Cremer et al.
1988). According to this law, asymmetry refiects the

Vertigo

272

~I

~O'

Fig.16.20. h-BPPV of the right ear. Schematic drawing of the right horizontal canal membranous duct illustrating the mechanism of
canalolithiasis in different head positions while supine. Rotation of the head of the patient around the longitudinal z-axis from a the
supine (nose up) to the right lateral, b right lateral to left lateral and c left lateral to right lateral positions while recumbent. Lower part
shows the induced horizontal eye movements, wh ich were most intense with ampullopetal stimuli (b versus c) and with maximal rotation angles of the head (a versus cl. The maximum slow-phase velocity (mean SD of n = number of positioning manoeuvres) was
54.6 13.6/s (n = 3) in a, 17.4 lO.4/s (n = 7) in band 176.2 13.00 /s (n = 9) in c. (From Strupp et al. 1995.)

dis charge properties of the vestibular nerve, which


can be increased above its resting level (between 70
and 100 Hz) much more (up to 500 Hz) than it can
be decreased. A mechanical explanation is unlikely
because the mass has to move the same distance in
both cases (Fig. 16.20) if the lumen of the membranous duct does not narrow within this segment. A
study on the geometry of the human horizontal (lateral) semicircular membranous canal (Curthoys and
Oman 1987) found no significant differences in the
diameter of this relevant segment of the horizontal
canal.
The diameter of the clot must be larger than the
diameter of the ampullofugal arm of the horizontal
canal membranous duct near its exit into the utricle.
Otherwise, the clot would move spontaneously out of
the horizontal canal into the utricular cavity when

the patient turns around in bed (Steddin and Brandt


1994a). This also explains why h-BPPV is non-fatiguing and why a single manoeuvre with barrel roll is
not successful in most cases. Baloh et al. (1993) proposed an alternative explanation for this amplitude
asymmetry. They suggest that the mass possibly has
a smaller distance to move in the canal when the
head is rotated toward the normal side than when
rotated toward the abnormal side.
That the duration of positioning nystagmus
(primary phase PI) is nearly twice as long for the
horizontal than for the posterior canal BPPV is due
to the greater amount of velo city storage in the horizontal canal ocular reflex than in the vertical canal
ocular reflex (Baloh et al. 1983). With step changes
in angular velocity the cupula of the posterior and
the horizontal canal returns to primary position

273

Benign paroxysmal positioning vertigo

PD

PI

~~~~L~~~~------u-I4oo
1 5 si

PI

~O'

1 5 si

1
I

40 '
40

Fig.16.21.

h-BPPVofthe right ear. Electronystagmographic recording (bitemporal horizontal eye movements, monocular vertical
movements) shows positioning nystagmus precipitated bya quick turn of the head from left lateral to right lateral and then vice versa
with the patient supine. (PI [immediate response], Pli [Iate-reversal response] = post-rotatory nystagmus land 11). (From Strupp et al.
1995.)

<?J ~..-..-_-----...-.-.._,...................."n.....

90

~_ _R_

____

1 5 Si

R
L

I
I

40

400

Fig.16.22. Electronystagmographic recording (horizontal eye movements) during positioning manoeuvres as in Fig. 16.21. After
successfulliberatory manoeuvres, no positioning nystagmus could be elicited. (From Strupp et al. 1995.)

with a time constant of about 7 s (Fernandez and


Goldberg 1971). The central velo city storage
mechanism prolongs this time constant through
central vestibular feedback pathways (Raphan et al.
1979). The velo city storage mechanism in the horizontal system is approximately twice that of the
vertical system, so that horizontal rotation-induced
nystagmus decays more slowly than vertical rota-

tion-induced nystagmus (Baloh et al. 1983). The


primary phase of positioning nystagmus (PI) may
be followed by areversal (PlI) comparable to
caloric irrigation of the horizontal semicircular
canal (Fig. 16.21). This reversal of positioning nystagmus is seen in about 50% of the patients, if the
head is rotated toward the affected ear and if the
initial nystagmus is intense with high slow-phase

Vertigo

274

Fig.16.23. h-BPPV, sehematic drawing of the left horizontal semicireular eanal with a free-floating elot in the dependent part of the
eanal (supine position, a). Rapid rolling onto the left (affeeted) ear eauses the elot to move toward the ampulla, whieh eauses an
ampullopetal defleetion with the typical BPPV attaek b. Rolling onto the right ear in supine position eauses the elot to move in the
opposite direetion in the eanal, whieh might induee ampullofugal defleetion of the eupula c. The anterior eanal and the eommon erus
are indieated by a and c, respeetively. (From Brandt and Steddin 1993.)

velo city (Baloh et al. 1993; De la Meilleure et al.


1996).
Transition of canalolithiasis to cupulolithiasis
It is theoretically conceivable that the free-floating

debris sometimes settle on and adhere to the cupula


(Steddin and Brandt 1994a,b; Welling and Barnes
1994). Apogeotropic and direction-changing positional nystagmus have been described as variants of
BPPV (Baloh et al. 1995; Nuti et al. 1996). Whereas
Nuti et al. (1996) hypothesised that the transformation of geotropic positional nystagmus into apogeotropic may be due to displacement of otoconia
from the anterior part of the canal to the posterior
part, we found evidence for transitions of
canalolithiasis to cupulolithiasis in the horizontal
semicircular canal (Steddin and Brandt 1996). We
described two patients in whom new positioning
manoeuvres for diagnostic and therapeutic purposes were tested. In the meantime, we have observed
this phenomenon in a few more patients suffering
from h-BPPV.
When the patients were in supine position (see
Fig. 16.24a-d), movement of the head from one side
to the other caused intense attacks of vertigo and
nystagmus. These were more pronounced (120 0 /s
versus 65/s nystagmus slow phase velo city) when
the head was turned toward the side of the affected
ear (Fig. 16.24c versus d). Vertigo and nystagmus
depended not only on the direction of head turn, but
also on the net angle of head rotation (Fig. 16.24b
versus d). The attacks lasted 50-70 s. The fast phase

of the purely horizontal nystagmus was in the direction of head movement (geotropic).
When the patients bent over from supine to the
"head-on-knees" position in combination with a
turn of the head toward the side of the unaffected
ear (Fig. 16.25b), vertigo and nystagmus were much
stronger, with the nystagmus directed to the side of
the affected ear. Maximum slow-phase velo city
reached 210 0 /s in Patient 1 and 370 0 /s in Patient 2. A
subsequent turn of the head to a face-down position
(Fig. 16.25c) eaused less intense nystagmus and vertigo of a shorter duration. The fast phases of the
nystagmus were in the same direction as the head
movement.
When the patients were returned to the supine
position, they experienced no vertigo or nystagmus.
On repetition of the initial head movements in Fig.
16.24, the patients showed completely different signs
and symptoms (Fig. 16.26): A 60 turn of the head to
either the right or left side caused purely horizontal
nystagmus toward the uppermost ear (apogeotropie), i.e. in the reverse direction of that resulting from
the identical manoeuvres in Fig. 16.24. The nystagmus was less intense and, unlike the effects in Fig.
16.24, depended only on the assumed head position
rather than the net angle of head rotation (Fig.
16.26b versus d). The time course of the nystagmus
also differed; it las ted for up to 3 minutes. Both
patients reported a much less distressing sensation
of rotational vertigo. Whereas in Patient 1 these features continued until the next day, in Patient 2 they
returned to the original behaviour depicted in Fig.
16.24 when retested after a 10-min break.

275

Benign paroxysmal positioning vertigo

right

( rl9ht eye )

..

'<t

rlght

20 sec

..

'<t

Fig.16.24. h-BPPV ofthe left ear ("primary" canalolithiasis; compare Fig. 16.9). Typical diagnostic positioning manoeuvres in supine
position with nose up (feft) are schematically depicted and the original (monocular horizontal) electronystagmographic recordings for
both eyes (right) of patient 1 are shown. The causative debris are depicted as a bfack circle within the drawing of a section of the left
horizontal canal; an open circle represents the final resting position of the debris.Arrows indicate the direction of debris motion. The
cupula is depicted as a thick fine within the ampulla of the canal; alignment of the cupula's plane is indicated by a dashed fine. a The
free-floating, heavy debris rest diametrically opposite to the cupula of the horizontal canal. b With a quick turn of the head 60' to the
right, the debris move away from the ampulla (ampullofugally) according to gravitational force, thus inducing an ampullofugal cupula
deflection with concomitant vertigo and horizontal nystagmus. cA subsequent head turn 120' to the left causes the debris to move
toward the ampulla (ampullopetally). Ampullopetal deflection of the cu pu la causes a more severe vertigo and horizontal positional
nystagmus. d If the head is now turned back by 120' (toward the right earl, ampullofugal stimulation causes vertigo and positional
nystagmus in the opposite direction; however, they are less intensive (although net angle of debris motion is the same as in cl. The
nystagmus in manoeuvres b-d is purely horizontal with fast phases beating toward the direction of the respective head movements.
(From Steddin and Brandt 1996.)

The change in direction of nystagmus and vertigo, time course, and intensity seen in the patients
shown in Figs. 16.24-26 can be best explained by the
different mechanisms of canalolithiasis and
cupulolithiasis. It remains unclear why a transition
from canalolithiasis to cupulolithiasis does not
occur more frequently in BPPV of the horizontal and
posterior semicircular canals. In any case, the headon-knees manoeuvre seems to be a way to repeatedly
provoke this transition in h-BPPV.
The important message for the clinician is that
positional nystagmus beating toward the uppermost

ear does not, contrary to previous belief, prove central positional nystagmus. It is explained by rare
cupulolithiasis of the horizontal canal.
Reversible ipsilateral caloric hypoexcitability

Pagnini et al. (1989) found that 5 of 15 and Baloh et


al. (1993) that 4 of l3 patients with h-BPPV had unilateral peripheral horizontal semicircular canal
paresis, which was detected by caloric irrigation. In
each case, the caloric hypoexcitability occurred on
the side undermost when the more prominent

276

Vertigo

( fight eye )

~------------20sec ------------~

left eye

Fig.16.25. Positional effects of a rapid bending-over manoeuvre (same patient as in Fig. 16.24) from supine with the head turned
right (a-b). This causes maximal vertigo and positional nystagmus due to a strong ampullopetal stimulation. Maximum slow-phase
velocity of positioning nystagmus reaches 370/5. With this manoeuvre the debris come dose to the cupula.lfthe head is then turned
to normal position c, vertigo and nystagmus are again induced in the same direction. The debris should now rest near the crista at the
base of the cupula. A subsequent head turn to the side of the affected ear caused no vertigo or nystagmus d. (From Steddin and
Brandt 1996.)

positioning vertigo and nystagmus were elicited. It is


surprising that De la Meilleure et al. (1996) saw only
a few cases of caloric hypoexcitability of the affected
or the unaffected ear in their large series of 63
patients. We share the experience of Pagnini et al.
(1989) and Baloh et al. (1993) that about 30% ofthe
patients presenting with h-BPPV have unilateral
caloric hypoexcitability. We documented the
reversibility of this condition following successful
liberatory manoeuvres (Fig. 16.27) and attribute the
caloric hypoexcitability in these patients to a partial
plugging of the horizontal semicircular canal
(Strupp et al. 1995).
There are convincing arguments that the caloric
response consists of two major components: convective and non-convective (Table 16.5). According to the
convective theory, originally proposed by Barany

(1906,1907), the thermal stimulus causes a temperature gradient across the end-organ, which results in
density changes in the endolymph and leads to a convective current when the fluid is acted upon by gravity. A strang argument for the convective mechanism
is the observation that caloric nystagmus becomes
inverted when the examined subject is turned from a
supine to a prane position (Coats and Smith 1967).
The 1983 Spacelab-1 experiments, however, raised
questions about Barany's theory. When caloric
nystagmus was elicited in a weightless environment,
direct evidence of the existence of a non-convective
component was found. Several mechanisms may
explain this component (Table 16.4):
1. thermal expansion of the labyrinthine fluids

(praposed by Scherer and Clarke (1985, 1987), a

277

Benign paroxysmal positioning vertigo

rlght

( rl ght eye)

~------------20sec------------~
right

le fteye

Fig. 16.26.

h-BPPV of the left ear ("secondary" cupulolithiasis; compare Fig. 16.8). The patient and the sequence of positioning
manoeuvres (a-d) are the same as in Figs. 16.24-25. ENG recordings indicate that the direction of the induced nystagmus in the challenging head positions is opposite to that in Fig. 16.24. The manoeuvres depicted here were performed directly after the extreme
bending-forward manoeuvres depicted in Fig. 16.25.The reversal of the direction of nystagmus is linked to areversal of cupular stimulation. With respect to the mechanism involved, this can only bE explained by a transition of canalolithiasis into cupulolithiasis. With
the debris fixed to the cupula, the semicircular canal changes from a sensor of rotational acceleration to a sensor of linear (gravity)
acceleration. The induced positional vertigo and nystagmus are less violent, but langer-lasting. (Fram Steddin and Brandt 1996.) ENG =
electronystagmographic.

strongly criticised hypo thesis (Hood 1989; Minor


and Goldberg 1990; Stahle 1990), and
2. direct thermal effects on vestibular hair cells,
synapses (Arechiga and CerbOn 1981), and afferents (Schermuly and Klinke 1985; Klinke 1992).
Vestibular hair cells exhibit motile responses to heat
(Zenner and Zimmermann 1991) as weIl as
increased intracellular free calcium ion concentration (Ohtani et al. 1993). Most authors believe that
the convective component accounts for 70 to 75% of
the caloric response (Gentine et al. 1991; Hood 1989;
Minor and Goldberg 1990).
The reversible ipsilateral caloric hypoexcitability
can be explained by functional plugging of the hori-

zontal canal, which would abolish the convective current. To eliminate the convective current, the dot (that
is assumed to have almost the same diameter as the
horizontal canal membranous duct) must be located
diametrically opposite or doser to the ampulla in the
horizontal canal, or both (Fig. 16.20c). This condition
is met when the head is slightly turned toward the
irrigated ear during caloric stimulation to facilitate
the ftow of water out of the extern al auditory canal.
Our observation of caloric hypoexcitability is also
supported by studies in squirrel monkeys: plugging of
the horizontal canal reduced caloric response (Paige
1985). Moreover, these results indicate that physiological caloric responses consist of a major convective
and a minor non-convective component.

right ear 44C

L
R

5s

right ear 30C

left ear 30C

5s

140
140
140
140

Fig.16.27. h-BPPV of the right ear. Caloric irrigation of the right and left ear before (top) and after (bottom) successfulliberatory
manoeuvres. The head of the supine patient was tilted upward 30 (conventional electronystagmography, recording bitemporal horizontal eye movements). Water at 44C and 30C was infused into the external meatus for 30 s at each temperature.The measured maximum horizontal slow-phase velocity (SPV, means SD of 12 measurements in three caloric stimulations) indicates a reversible caloric
hypoexcitability of the right horizontal canal (HC). For water at 44C before versus after liberatory manoeuvres: right HC SPV 5.4
1.3/s versus 36.0 8.3/s, left HC SPV 36.2 3.9/s versus 32.7 2.rl s. For 30C before versus after liberatory manoeuvres: right HC
SPV 7.3 0.8Is versus 26.3 3.0Is, left HC SPV 28.5 4.0Is versus 28.2 4.1 /s. (From Strupp et al. 1995.)

Management
Several features of h-BPPV remain unclear and are
still a subject of speculation. For instance, why does
canalolithiasis of the horizontal semicircular canal
occur despite the fact that debris leave the canal on a

simple head or body tilt from one side to the other,


as is often performed while lying in bed? We suspect
that the following two conditions must be fulfilled
for the debris to remain in the canal (Steddin and
Brandt 1996):

279

Benign paroxysmal positioning vertigo

1. The diameter of the congealed debris must be

greater than that of a bottleneck-like narrowing


of the distal branch of the canal.
2. The configuration of the debris, which congeal in
the canal, must be so stable that the dot does not
break into pieces small enough to pass the
bottleneck.
If the fatiguability of symptoms on repetitive testing
is explained by transient dissolving of the debris,
then non-fatiguability, which is frequently seen in
h-BPPV, may prove the above assumption.
Consequently, as long as there is no fatigue of vertigo
and nystagmus in h-BPPV, manoeuvres intended to
sluice the debris out of the canal should have minor
success. In fact, the liberatory manoeuvres that
Lempert (1994; Lempert and Tiel-Wilck 1996) and
Baloh (1994) proposed were each based on only two
patients. Lempert (1994) described a single 270
"barbecue rotation" toward the unaffected side,
which was performed in rapid steps of 90 at 30second intervals. Baloh (1994) suggested a 360 rotation around the yaw axis in four quick steps of 90
with the initial motion toward the healthy side; each
position was held for about 1 minute.
Vannucchi et al. (1997) compared the therapeutic
results obtained by maintaining a prolonged position on the healthy side (35 patients) with repetitive
head shaking in a supine position (24 patients) and
no therapy (15 patients). More than 90% of the
patients treated with prolonged position recovered
within 3 days, although 6 of 35 patients subsequently
developed p-BPPV (which responded successfully to
repositioning manoeuvres). The rationale was
obviously that the "heavy partides" in the nonampullary arm of the horizontal canal gradually
moved out when the patient maintained a prolonged
position on the side of the unaffected ear. This
manoeuvre was not effective if performed only for a
duration of 10-20 mins; it was intended to be maintained for up to 12 hours. In those who failed to
respond, the Brandt-Daroff exercises (Brandt and
Daroff 1980) were performed and within a matter of
days led to full recovery. This agrees with the report
of Baloh et al. (1993) and our own experience with
such patients. Nuti et al. (1998) compared the effects
of"barbecue rotation" versus forced prolonged position in a large group of 92 patients with h-BPPV.
Forced prolonged position was successful in more
than 70%, while the "barbecue rotation" was slightly
less successful but had more immediate results.
Thus, for the time being, we first try "barbecue
rotation" and propose that prolonged bedrest with
the head turned toward the unaffected ear
(Vannucchi et al. 1997) be maintained for up to 12
hours. If this is still unsuccessful after 2 days, we
advise the patients to perform the Brandt-Daroff

Table 16.5.

Possible mechanisms of physiological caloric response

Convective component
(dependency on gravity/
head position, elimination
by plugging of the HC
(Paige 1985))

Mechanisms

References

Temperature gradient,
resulting in density
changes in the
endolymph, leading to
convection current or
"hydrostatic model"

Barany (1906, 1907)


Steinhausen (1921)
Coats and Smith
(1967)

Nonconvective component
(not dependent on
gravity/head position*)

Gentine et al. (1991)


Hood (1989)

(1) Fluid expansion


("direct volume
displacement")
("hydrostatic
pressure change")
(2) Direct thermal
effect on:
Vestibular ha ir cell
(intracellular Ca2+,
length of hair cell)
Synapse
Nerve

Scherer and C1arke


(1985)
Baumgarten et al.
(1985)

Barteis (1911)
Ohtani et al. (1993)
Zenner and
Zimmerman (1991)
Arechia and Cerb6n
(1981)
Schermulyand
Klinke (1985)

* Animal studies have described a position-dependent modulation of the


non-convective component under conditions of 1 9 after the horizontal
canal (HC) was plugged (Paige 1985; Minor and Goldberg 1990). This could
result from a position-dependent, possibly otolith-mediated, modulation
of the horizontal vestibulo-ocular reflex (Minor and Goldberg 1990). (From
Strupp et al. 1995.)

exercises (Fig.16.15; p. 265). Both physical therapies


can be performed at horne and do not require the
presence of a physical therapist.

Anterior semicircular canal BPPV


(a-BPPV)
BPPV of the anterior semicircular canal is a rare
variant that sometimes occurs inadvertently as a
transition from typical p-BPPV to a-BPPV while
patients are undergoing physical liberatory manoeuvres (Herdman and Tusa 1996). There is little
documentation of this type of BPPV in the literature.
According to the effects of ampullofugal posterior
canal stimulation (Fig. 16.28), rotatory vertigo in an
a-BPPV attack is associated with

positioning nystagmus that is downbeating and


torsional toward the affected uppermost ear
(Table 16.6)

Vertigo

280
Table 16.6. Summary of oculomotor examination in vertical or horizontal semicircular (anal BPPV.
(anal
involvement

Hallpike-Dix

Posterior

Upbeating and
torsional'
Downbeating and
torsional
Horizontal

Anterior
Horizontal

Nystagmus
Hallpike-Dix
reversal

Return to sitling

Downbeating

Downbeating

Upbeating

Upbeating

Horizontal
(opposite
direction)

Horizontal

'Fast phase of nystagmus with superior pole of eye beating toward the
downside ear. (From Herdman and Tusa 1996.)

Fig. 16.28. Anterior semicircular canal BPPV of the left


labyrinth as seen from the occiput: with the head upright the dot
should rest on the cupula of the anterior canal (Jeft) . A rapid 145 0
head-trunk tilt toward the unaffected ear into a head-hanging
position results in ampullofugal stimulation because the elot
gravitates toward the common crus of anterior and posterior
canals (right).

latency, duration, and directional reversal


(upbeat and torsional toward the unaffected ear)
on uprighting the patient, effects that are similar
to that of typical p- BPPV.

An effective positioning manoeuvre to assess the


diagnosis of a-BPPV is

the Dix and Hallpike manoeuvre into a headhanging position with the head turned toward
the unaffected ear (Fig. 16.29,1), which results in
ampullofugal stimultion
rapid uprighting of the patient from this position (Fig. 16.29,2) which results in ampullopetal
stimulation

Thus, a-BPPV is also elicited when the affected ear is


uppermost (Brandt et al. 1994). When the patient's
head is in the normal upright position (Fig. 16.28),
the clot should rest on the ampulla of the anterior
canal; this theoretically causes cupulolithiasis.

Fig.16.29. a-BPPV of the left labyrinth: schematic drawing of


precipitatory positioning manoeuvres. A 145 0 head tilt in the
plane of the anterior semicircular canal (head rota ted 45 0 toward
the unaffected earl will cause ampullofugal stimulation of the
anterior canal (1) . Rapid uprighting will result in ampullopetal
stimulation (2).

No controlled studies are available on the efficacy


of different physical manoeuvres for a-BPPV. In
theory, the appropriate liberatory manoeuvres are
similar to those of p-BPPV of the contralateral ear
(Fig. 16.16, p. 266), since the spatial orientation of
the anterior canal of one ear is alm ost in the same
plane as that of the posterior canal of the other ear.
Brandt and Daroff exercises were effective in the few
patients with a-BPPV whom we treated outside the
framework of a controlled study and were superior
to single liberatory manoeuvres.

References
Agus G, Puxeddu R, Demontis GP, Puxeddu P (1995) Atypical
"reversed" paroxysmal positioning nystagmus in benign paroxysmai positional vertigo. Acta Otolaryngol (Stockh) Suppl
520:143-147
Allen G, Fernandez C (1960) Experimental observations in postural nystagmus: extensive lesions in posterior vermis of the
cerebellum. Acta Otolaryngol (Stockh) 51:2-14
Anastasopoulos D, Lempert T, Gianna C, Gresty MA, Bronstein
AM (1997) Horizontal otolith-ocular responses to lateral translation in benign paroxysmal positional vertigo. Acta
Otolaryngol (Stockh) 117:468-471
Andaz C, Whittet HB, Ludman H (1993) An unusual cause of
benign paroxysmal position al vertigo. J Laryngol Otol
107:1153-1154
Anthony PF (1991) Partitioning of the labyrinth: application in
benign paroxysmal positional vertigo. Am J OtoI12:388-393
Anthony PF (1993) Partitioning the labyrinth for benign paroxysmal positional vertigo: clinical and histologic findings. Am J
OtoI14:334-342
Arai Y, Henn V, Boehmer A, Suzuki J (1989) How could canalpluggings result in intensive direction changing type of positional nystagmus? Acta Otolaryngol (Stockh) Suppl
468:159-164
Arechiga H, Cerbbn J (1981) The influence of temperature and
deuterium oxide on the spontaneous activity of crayfish
motoneurons. Comp Biochem Physiol A 69:631-636

Benign paroxysmal positioning vertigo


Baloh RW (1994) Reply to the letter by Lempert: Horizontal
benign positional vertigo. Neurology 44:2214
Baloh RW, Sakala SM, Honrubia V (1979) Benign paroxysmal
positional nystagmus. Am J Otolaryngoll:I-5
Baloh RW, Hornrubia V, Jacobson K (1987) Benign positional vertigo. Neurology 37:371-378
Baloh RW, Jacobson K, Honrubia V (1993) Horizontal semicircular
canal variant of benign positional vertigo. Neurology
43:2542-2549
Baloh RW, Richman L, Yee RD, Honrubia V (1983) The dynamics
of vertical eye movements in normal human subjects. Aviat
Space Environ Med 54:32-38
Baloh RW, Yue Q, Jacobson KM, Honrubia V (1995) Persistent
direction-changing positional nystagmus: Another variant of
benign positional vertigo? Neurology 45: 1297 -130 I
Baniny R (1906) Untersuchungen ber den vom Vestibularapparat
des Ohres reflektorisch ausgelsten rhythmischen Nystagmus
und seine Begleiterscheinungen. Mschr Ohrenheilkd
Laryngorhinol 40: 191-297
Baniny R (1907) Physiologie und Pathologie (Funktionsprfung)
des Bogengangapparates beim Menschen. Franz Deuticke,
Leipzig
Barany R (1921) Diagnose von Krankheitserscheinungen im
Bereiche des Otolithenapparates. Acta Otolaryngol (Stockh)
2:334-437
Barber HO (1964) Positional nystagmus especially after head
injury. Laryngoscope 73:391-944
Barteis M (1911) Funktionelle Prfung des Vestibularapparates.
Ber Dtsch Otol Ges 20:214-215
Baumgarten von R, Boehmer G, Brenske A, Reiser M (1985) A
non-thermoconvective mechanism generating caloric nystagmus -similarities between pressure-induced and caloric nystagmus in the pigeon. In: Proceedings of the congress of the
Barimy Society. Ann Arbor, Michigan
Black FO, Nashner LM (1984) Vestibulo-spinal control differs in
patients with reduced versus dis tor ted vestibular function. Acta
Otolaryngol (Stockh) SuppI406:110-114
Black FO, Wall III C, Nashner LM (1983) Effects of visual and support surface orientation references upon postural control in
vestibular deficient subjects. Acta Otolaryngol (Stockh)
95:199-210
Bloom J, Katsarkas A (1989) Paroxysmal positional vertigo in the
elderly. J OtolaryngoI18:96-98
Brandt Th (1991) Positional and positioning vertigo and nystagmus (review article). J Neurol Sci 95:2-28
Brandt Th, Daroff RB (1980) Physical therapy for benign paroxysmal positional vertigo. Arch Otolaryngol 106:484-485
Brandt Th, Dieterich M (1993) Vestibular falls. J Vestib Res 3:3-14
Brandt Th, Steddin S (I992) Reply to the letter by Gordon: Benign
paroxysmal positional vertigo (BPPV) or bubble provoked
positional vertigo? J Neurol Sci II 1:231-233
Brandt Th, Steddin S (1993) Current view of the mechanism of
benign paroxysmal positioning vertigo: Cupulolithiasis or
canalolithiasis? J Vestib Res 3:373-382
Brandt Th, Steddin S, Daroff RB (1994) Therapy for benign
paroxysmal positioning vertigo, revisited. Neurology
44:796-800
Bchele W, Brandt Th (1979) Vestibulo-spinal ataxia in benign
paroxysmal positional vertigo. Agressologie 20:221-222
Bchele W, Brandt Th (1986) Benign paroxysmal positional vertigo and posture. In: Bles W, Brandt Th (eds), Disorders of posture and gait. Elsevier, Amsterdam, pp 141-156
Bchele W, Brandt Th (1988) Vestibular neuritis-a horizontal
semieircular canal paresis? Adv Oto-Rhino-Laryngol
42:157-161
Coats AC, Smith SY (1967) Body position and the intensity of
caloric nystagmus. Acta Otolaryngol (Stockh) 63:515-532
Cohen B, Tokumasu K, Goto K (1966) Semicircular canal nerve,
eye and head movements: the effect of changes in initial eye

281
and head position on the plane of the induced movement. Arch
Ophthalmol (Chicago) 76:523-531
Cremer PD, Henderson CJ, Curthoys IS, Halmagyi GM (1988)
Horizontal vestibulo-ocular reflexes in humans with only one
horizontal semicircular canal. Adv Otorhinolaryngol
42:180-184
Curthoys JS, Oman CM (1987) Dimensions of the horizontal semicircular duct ampulla and utricle in human. Acta Otolaryngol
(Stockh) 103:254-261
De la Meilleure G, Dehaene I, Depondt M, Damman W, Crevits L,
Vanhooren G (1996) Benign paroxysmal positional vertigo of
the horizontal canal. J Neurol Neurosurg Psychiatry 60:68-71
Dichgans J, Mauritz KH, Allum JHJ, Brandt Th (1975) Postural
sway in normals and atactic patients: analysis of the stabilizing
and destabilizing effects of vision. Agressologie 17C: 15-24
Dingle AF, Hawthorne MR, Kumar BU (1992) Fenestration and
occlusion of the posterior semicircular canal for benign positional vertigo. Clin OtolaryngoI17:300-302
Dix R, Hallpike CS (1952) The pathology, symptomatology and
diagnosis of certain common disorders of the vestibular
system.Ann Otol Rhinol LaryngoI6:987-1016
Epley JM (1980a) New dimensions of benign paroxysmal
positional vertigo. J OtolaryngoI8:151-158
Epley JM (l980b) New dimensions of benign paroxysmal positional vertigo. Otolaryngol Head Neck Surg 88:599-605
Epley JM (1992) The canalith repositioning procedure: for treatment of benign paroxysmal positional vertigo. Otolaryngol
Head Neck Surg 107:399-404
Epley JM (1995) Positional vertigo related to semicircular
canalithiasis. Otolaryngol Head Neck Surg 112:154-161
Fernadez C, Goldberg JM (1971) Physiology of peripheral neurons
innervating semi-circular canals of the squirrel monkey. II.
Response to sinusoidal stimulation and dynamics of peripheral
vestibular system. J NeurophysioI34:661-675
Fetter M, Sievering F (1995) Three-dimensional eye movement
analysis in benign paroxysmal positioning vertigo and nystagmus. Acta Otolaryngol (Stockh) 115:353-357
Fluur E (1974) Positional and positioning nystagmus as a result of
utriculo-cupular integration. Acta Otolaryngol (Stockh)
78:19-27
Froehling DA, Silverstein MD, Mohr DN, Beatty CW, Offord KP,
Ballard DJ (1991) Benign positional vertigo: incidence and
prognosis in a population-based study in Olmsted county,
Minnesota. Mayo Clin Proc 66:596-601
Gacek RR (1978) Further observations on posterior ampullary
nerve transection for positional vertigo. Ann Otol Rhinol
LaryngoI87:300-306
Gacek RR (1984) Cupulolithiasis and posterior ampullary nerve
transection. Ann Otol Rhinol Laryngol (SuppI1l2) 93:25-29
Gacek RR (1995) Technique and results of singular neurectomy
for the management of benign paroxsymal positional vertigo.
Acta Otolaryngol (Stockh) 115:154-157
Gentine A, Eichhorn JL, Koop C, Conraux C (1991) Modelling the
action of caloric stimulation of the vestibule. IV. The global
mechanical model. Acta Otolaryngol (Stockh) 111:633-638
Gordon AG (1992) Benign paroxysmal positional vertigo (BPPV)
or bubble provoked positional vertigo? J Neurol Sei
111 :229-230
Gordon N (1954) Post-traumatic vertigo, with special reference to
positional nystagmus. Lancet i: 1216-1218
Graf W, Ezure K (1986) Morphology of vertical canal related second order vestibular neurons in the cat. Exp Brain Res 63:35-48
Graf W, McCrea RA, Baker R (1983) Morphology of posterior
canal-related secondary vestibular neurons in rabbit and cat.
Exp Brain Res 52:125-138
Gregorius FK, Grandal PH, Baloh RW (1976) Positional vertigo
with cerebellar astrocytoma. Surg NeuroI6:283-286
Gussen R (1980) Saccule otoconia displacement into cochlea in
cochleosaccular degeneration. Arch Otolaryngol 106: 161-166

282
Gyo K (1988) Benign paroxysmal positioning vertigo as a complication of postoperative bedrest. Laryngoscope 98:332-333
Hall SF, Ruby RR, McClure JA (1979) The mechanics of benign
paroxysmal vertigo. J OtolaryngoI8:151-158
Halmagyi GM, Curthoys IS, Cremer PD et al (1990) The human
horizontal vestibulo-ocular reflex in response to highacceleration stimulation before and after unilateral vestibular
neurectomy. Exp Brain Res 81:479-490
Harbert F (1970) Benign paroxysmal positional nystagmus. Arch
OpthamoI84:298-302
Harvey SA, Hain TC, Adamiec LC (1994) Modified liberatory
manoeuvre: effective treatment for benign paroxysmal positional vertigo. Laryngoscope 104:1206-1212
Hasegawa T (1933) Die Vernderung der labyrinthren Reflexe bei
zentrifugierten Meerschweinchen. Pflgers Arch 232:454-465
Husler R, Pampurik JC (1989) Die chirurgische und die physiotherapeutische Behandlung des benignen paroxysmalen
Lagerungsschwindels. Laryngol-Rhinol-OtoI68:342-346
Hawthorne M, EI-Naggar M (1994) Fenestration and occlusion of
posterior semicircular canal for patients with intractable benign
paroxysmal positional vertigo. J Laryngol Otoll08:935-939
Herdman SJ (1990) Treatment of benign paroxysmal positional
vertigo. Phys Ther 70:381-387
Herdman SJ, Tusa RJ (1996) Complications of the canalith repositioning procedure. 122:281-286
Herdman SJ, Tusa RJ, Zee DS, Proctor LR, Mattox DE (1993) Single
treatment approaches to benign paroxysmal positional vertigo.
Arch Otolaryngol Head Neck Surg ll9:450-454
Hood JD (1989) Evidence of direct thermal action upon the
vestibular receptors in the caloric test. Acta Otolaryngol
(Stockh) 107:161-165
Igarashi M, Nagaba M (1968) Vestibular end-organ damage in
squirrel monkeys after exposure to intensive linear acceleration. In: NASA SP: third symposium on the role of the vestibular organs in space exploration. pp 152:63-71
Johnson L-G, Hawkins JE (1972) Sensory and neural degeneration
with ageing, as seen in microdissections of the human inner
ear.Ann OtoI81:179-193
Kanayama R, Bronstein AM, Gresty MA, Brookes GB (1995)
Vertical and torsional VOR in posterior canal occlusion. Acta
Otolaryngol (Stockh) SuppI520:362-365
Kartusch JM, Sargent EW (1995) Posterior semicircular canal
occlusion for benign paroxysmal positional vertigo -C0 2 Iaserassisted technique: preliminary results. Laryngoscope
105:268-273
Katsarkas A (1991) Electronystagmographic (ENG) findings in
paroxysmal positional vertigo (PPV) as a sign of vestibular
dysfunction. Acta Otolaryngol (Stockh) 111:193-200
Katsarkas A, Kirkham Th (1978) Paroxysmal positional vertigo as
a complication of postoperative bedrest. Laryngoscope
98:332-333
Katsarkas A, Outerbridge JS (1983) Nystagmus of paroxysmal
positional vertigo.Ann Otol Rhinol LaryngoI92:146-150
Klinke R (1992) Thermal effects on the vestibular hair cell
synapse have to be considered for the explanation of caloric
nystagmus. Acta Otolaryngol (Stockh) 112:574-576
Kveton JF, Kashgarian M (1994) Particulate matter within the
membranous labyrinth: pathologic or normal? Am J Otol
15:173-176
Lanska DJ, Remler B (1997) Benign paroxysmal positioning vertigo: classic descriptions, origins of the provocative positioning
technique, and conceptual developments. Neurology
48:ll67-ll77
Lempert T (1994) Horizontal benign positional vertigo (letter).
Neurology 44:2213-2214
Lempert T, Tiel-Wilck K (1996) A positional manoeuvre for treatment of horizontal-canal benign positional vertigo.
Laryngoscope 106:476-478
Lempert T, Wolsley C, Davies R, Gresty MA, Bronstein AM (1996)

Vertigo
Curing benign positional vertigo in a 3D flight simulator.
Lancet 347:1192
Leuwer RM, Westhofen M (1996) Surgical anatomy of the singular
nerve. Acta Otolaryngol (Stockh) ll6:576-580
Li JC (1995) Mastoid oscillation: a critical factor for success in the
canalith repositioning procedure. Otolaryngol Head Neck Surg
112:670-675
Lindsay JR, Hemenway WG (1956) Postural vertigo due to unilateral sudden partialloss of vestibular function. Ann Otol Rhinol
Laryngol 65:692-708
Longridge NS, Barber HO (1978) Bilateral paroxysmal positioning
nystagmus. Can J Otol 7:395-400
Markham CH, Diamond SG, Juichi I (1987) Utricular dysfunction
in benign paroxysmal positional vertigo. In: Graham MD,
Kemink L (eds) The vestibular system: neurophysiologic and
clinical research. Raven Press, New York, pp 275-283
McClure JA (1985) Horizontal canal BPPV. J OtolaryngoI14:30-35
McClure JA (1988) Functional basis for horizontal canal BPPV. In:
Barber HO, Sharpe JA (eds) Vestibular dis orders. Year Book
Medical Publishers, Chicago, pp 233-238
Minor LB, Goldberg JM (1990) Influence of static head position on
the horizontal nystagmus evoked by caloric, rotational and
optokinetic stimulation in the squirrel monkey. Exp Brain Res
82:1-13
Mizukoshi K, Watanabe Y, Shojaku H, Okubo J, Watanabe I (1988)
Epidemiological studies on benign paroxysmal positional vertigo in Japan. Acta Otolaryngol (Stockh) SuppI447:67-72
Money KE, Myles WS (1974) Heavy water nystagmus and effects of
alcohol. Nature 247:404-405
Money KE, Myles WS, Hoffert BM (1974) The mechanism of positional alcohol nystagmus. Can J Otolaryngol 3:302-313
Moriarty B, Rutka J, Hawke M (1992) The incidence and distribution of cupular deposits in the labyrinth. Laryngoscope
102:56-59
Naganuma H, Kohut RI, Tokumasu K, Okamoto M, Fujino A, Arai
M (1996) Basophilic deposits on the cupula: preliminary findings describing the problems involved in studies regarding the
incidence of basophilic deposits in the cupula. Acta
Otolaryngol (Stockh) SuppI524:9-15
Nomura Y, Ooki S, Kukita N, Young YH (1995) Laser labyrinthectomy.Acta Otolaryngol (Stockh) ll5:158-161
Norre ME (1995) Reliability of examination data in the diagnosis
ofbenign paroxysmal positional vertigo.Am J OtoI16:806-81O
Nuti D, Vannucchi P, Pagnini P (1996) Benign paroxysmal positional vertigo of the horizontal canal: a form of canalolithiasis
with variable clinical features. J Vestib Res 6:173-184
Nuti D,Agus G, Barbieri M-T, Passali D (1998) The management
of horizontal-canal paroxysmal positional vertigo. Acta
Otolaryngol (Stockh) 118:455-460
Ohtani M, Yamashita T, Amano H, Kubo N, Kumazawa T (1993)
Thermal influence on intracellular calcium concentration in
vestibular hair cells isolated from the guinea pig. Acta
Otolaryngol Suppl (Stockh) 500:46-49
Pace-Balzan A, Rutka JA (1991) Non-ampullary plugging of the
posterior semicircular canal for benign paroxysmal positional
vertigo. J Laryngol OtoI105:901-906
Pagnini P, Nuti D, Vannucchi P (1989) Benign paroxysmal vertigo
of the horizontal canal. ORL J Otorhinolaryngol Relat Spec
51:161-170
Paige GD (1985) Caloric responses after horizontal canal inactivation. Acta Otolaryngol (Stockh) 100:321-327
Parnes LS, McClure JA (1990) Posterior semicircular canal occlusion for intractable benign paroxysmal positional vertigo. Ann
Otol Rhinol LaryngoI99:330-334
Parnes LS, McClure JA (1991) Posterior semicircular canal occlusion in the normal hearing ear. Otolaryngol Head Neck Surg
104:52-57
Parnes LS, McClure JA, (1992) Free-floating endolymph particles.
Laryngoscope 102:988-992

Benign paroxysmal positioning vertigo


Raphan T, Matsuo V, Cohen B (1979) Velo city storage in the
vestibulo-ocular reflex arc. Exp Brain Res 35:229-248
Rietz R, Troia BW, Yonkers AI, Norris TW (1987) Glycerol-induced
positional nystagmus in human beings. Otolaryngol Head Neck
Surg 97:282-287
Sakata E, Ohtsu K, Shimura H, Sakai S (1987) Positional nystagmus of benign paroxysmal type (BPPV) due to cerebellar
vermis lesions: pseudo-BPPV. Auris Nasus Larynx (Tokyo)
14:17-21
Sakata E, Ohtsu K, Itoh Y (1991) Positional nystagmus ofbenign
paroxysmal type (BPPN) due to cerebellar vermis lesions.
Pseudo-BPPN. Acta Otolaryngol (Stockh) SuppI481:254-257
Scherer H, Clarke AH (1985) The caloric vestibular reaction in
space. Acta Otolaryngol (Stockh) 100:328-336
Scherer H, Clarke AH (1987) Thermal stimulation of the vestibular labyrinth during orbital flight. Arch Otorhinolaryngol
244: 159-166
Schermuly L, Klinke R (1985) Change of characteristic frequency
of pigeon in primary auditory afferents with temperature. J
Comp PhysioI156:209-211
Schuknecht HF (1962) Positional vertigo. Clinical and experimental observations. Trans Am Acad Ophthalmol Otolaryngol
66:319-331
Schuknecht HF (1969) Cupulolithiasis. Arch Otolaryngol
90:765-778
Schuknecht HF, Ruby RRF (1973) Cupulolithiasis. Adv Oto-RhinoLaryngol 22:434-443
Semont A, Freyss G, Vitte E (1988) Curing the BPPV with a liberatory manoeuvre. Adv OtolrhinolaryngoI42:290-293
Serafini G, Palmierei AMR, Simoncelli C (1996) Benign paroxysmal positional vertigo of posterior semicircular canal: results in
160 cases treated with Semont's manoeuvre. Ann Otol Rhinol
LaryngoI105:770-775
Smouha EE (1997) Time course of recovery after Epley manoeuvres for benign paroxysmal positional vertigo.
Laryngoscope 107:187-191
Stahle J (1990) Controversies on the caloric response. From
Baniny's theory to studies in microgravity. Acta Otolaryngol
(Stockh) 109:162-167
Steddin S, Brandt Th (1994a) Benigner paroxysmaler
Lagerungsschwindel: Differentialdiagnose der posterioren, horizontaen und anterioren Kanalolithiasis. Nervenarzt 65:505-510
Steddin S, Brandt Th (1994b) Unilateral mimicking bilateral
benign paroxysmal positioning vertigo. Arch Otolaryngol Head
Neck Surg 120: 1339-1341
Steddin S, Brandt T (1996) Horizontal canal benign paroxysmal

283
positioning vertigo (h-BPPV): translation of canalolithiasis to
cupulolithiasis. Ann NeuroI40:918-922
Steenerson RL, Cronin GW (1996) Comparison of the canalith
repositioning procedure and vestibular habituation training in
forty patients with benign paroxysmal positional vertigo.
Otolaryngol Head Neck Surg 114:61-64
Steinhausen W (1921) ber den experimentellen Nachweis der
Endolymphbewegung
im
Bogengangsapparat
des
Ohrlabyrinths bei adquater und kalorischer Reizung. Pflgers
Arch 187:47-74
Strupp M, Brandt Th, Steddin S (1995) Horizontal canal benign
paroxysmal positioning vertigo: Reversible ipsilateral caloric
hypoexcitability caused by canalolithiasis? Neurology
45:2072-2076
Suzuki M, Kadir A, Hayashi N, Takamoto M (1996) Functional
model of benign paroxysmal positional vertigo using an isolated
frog semicircular canal. J Vestib Res 6:121-125
Szentagothai I (1950) The elementary vestibulo ocular reflex arc. J
NeurophysioI13:395-407
Troost BT, Patton JM (1992) Exereise therapy for positional vertigo. Neurology 42: 1441-1444
Vannucchi P, Giaonnoni B, Pagnini P (1997) Treatment of horizontal semieircular canal benign paroxysmal positional vertigo. J
Vestib Res 7:1-6
Vogel K (1950) Zur Entstehung des peripheren Lagenystagmus.
Arch Ohr Heilk u Z Hals-usw Heilk157:89-98
Voorhees RL (1989) The role of dynamic posturography in
neurootologic diagnosis. Laryngoscope 99:995
Vyslonzil E (1963) ber eine umschriebene Ansammlung von
Otokonien in hinteren hutigen Bogengngen. Monatschr
Ohrenheilkd 97:63
Watson P, Barber HO, Peck H, Terbrugge K (1981) Positional vertigo and nystagmus of central origin. Can J Neurol Sei
8:133-137
Welling DB, Barnes DE (1994) Particle repositioning manoeuvre
for benign paroxysmal positional vertigo. Laryngoscope
104:946-949
Welling DB, Parnes LS, O'Brien B, Bakaletz LO, Brackman DE,
Hinojosa R (1997) Particulate matter in the posterior semieircular canal. Laryngoscope 107:90-94
Yagi T, Ushio K (1995) Nystagmus in benign paroxysmal positional
vertigo: a three-component analysis. Acta Otolaryngol (Stockh)
520:238-240
Zenner HP, Zimmermann U (1991) Motile responses of vestibular
hair cells following caloric, electrical or chemical stimuli. Acta
Otolaryngol (Stockh) 111:291-297

Positional nystagmus/vertigo with


specific gravity differential between
cupula and endolymph (buoyancy
hypothesis )
Transient positional nystagmus has been repeatedly
observed following the ingestion of water- and lipidsoluble molecules with specific gravities differing
from that of endolymph, such as alcohol or "heavy
water". The semicircular canals selectively transduce
angular velocity and head acceleration, and under
normal circumstances are insensitive to gravitational
orientation and linear acceleration. A major reason
for this insensitivity to head orientation in space is
that the cupula and endolymph have the same
specific gravity of 1.0087 relative to water (the sensory
hair cells are embedded in the cupula which is
housed in the ampulla of the canals). The neutral
buoyancy of the cupula in the endolymph prevents
any out-of-balance forces when linear accelerations
are applied.
Jf there is a considerable specific gravity differential between cupula and endolymph, the semicircular canals will become sensitive to changes in head
position within the gravitational field, resulting in
positional rotatory vertigo and nystagmus. The
direction of both nystagmus and vertigo shou ld be
dependenton
1. the particular head position (according to the
different planes of the horizontal and vertical
semicircular canals) and
2. on wh ether the specific gravity of the cupula is
greater or less than that of the endolymph.

Thus, nystagmus should change direction with


either head position right lateral or left lateral and
beat
1. toward the undermost ear when the cupula is
"lighter" than endolymph or
2. toward the uppermost ear when the cupula is
"heavier" than endolymph.

This hypothesis - also called the buoyancy hypothesis


- requires that ingested water-soluble compounds of
different specific gravity diffuse at different speeds
into cupula and endolymph, thus causing a transient

density gradient. Experiments conducted with


ethanol, deuterium oxide, and glycerol induced positional nystagmus, which is consistent with the
hypo thesis.
The pathophysiological concept of a gravity differential between cupula and endolymph, which
causes a positional vertigo/nystagmus (alcohol,
glycerol, heavy water, macroglobulinaemia), leaves
several questions open (Brandt 1990). For example,
why is positional alcohol nystagmus maintained for
minutes to hours in the precipitating head position?
Is it because the gravity-dependent deflection force
is greater than the physiological restoring force? Tbis
explanation is consistent with a constant deflection
of the cupula, which would provide continuous
mechanical stimulation of the hair cells. The gravitational force would be balanced by the elastic restoring force of the cupula tissue at some deflected
position (Fig. 17.1). The normal time constant (7 s
time constant of the cupula; cupulogram) seems to
result from the endolymph viscosity opposing the
elastic restoring force that is intrinsic to the cupula
when the stimulus is removed. However, it seems
strange that the hair cells fail to adapt, which is
implied by this long-lasting response.
The common view that benign paroxysmal positioning vertigo (see BPPV, p. 257) is induced by the
same mechanism (Schuknecht 1969; Rietz et al. 1987)
cannot be correct, however, for several reasons. First,
BPPV is a positioning rather than positional vertigol
nystagmus, because it is induced only by rapid
changes in head position. Second, BPPV subsides
after 10-60 sand ultimately abates even when the
precipitating position is maintained. Third, symptoms of vertical and horizontal canal BPPV are compatible with a free-floating dot (canalolithiasis, see
Chap. 16, p. 259) rather than a "heavy cupula" (cupulolithiasis).
If the concept of cupulolithiasis, "heavy cupula:' is
valid, two findings remain undear. Why do patients
with BPPV not suffer from additional positional
vertigo/nystagmus? Why do compounds with differing specific weights (such as alcohol and "heavy

285

Vertigo

286

water") induce positional but not positioning nystagmus with rapid changes in head position? If they
do, then for alcohol this positioning nystagmus
should beat toward the same direction as position al
nystagmus du ring the resorption phase (PA r,
when the cupula is relatively lighter) but toward the
opposite direction du ring the reduction phase (PAN
II, when the cupula is relatively heavier).
There is little mention of possible effects of the
various molecules on endolymph viscosity, neural
activity, and adaptation (Zucca et al. 1995; Takumida
et aJ. 1995), and one may wonder if this has any relevance to the difficulty of reconciling theory and clinical finding ? Depending on the diameter of the
semicircular canal, lower endolymph viscosity
should produce a faster response (tending to sense
acceleration rather than velocity) and a shorter time
constant of the system. Higher viscosities should
lead to reduced gain and a longer time constant for
decay of the output, unless the adaptation of the sensory cells to a constant stimulus has a significant
effect. Further experiments are needed to clarify
these discrepancies between theory and clinical
manifestation. As fascinating as positional nystagmus and vertigo are, if they are caused by the buoyancy mechanism, the clinical relevance of the four
conditions is minimal:

positional alcohol vertigo/nystagmus (PAN)


positional "heavy water" nystagmus
positional glycerol nystagmus
positional nystagmus with macroglobulinaemia

In the differential diagnosis of positional nystagmus


(p. 247), it is helpful for the clinician to know that in
outpatients, who present after alcohol excess the
night before, PA beats toward the uppermost ear
(alcohol reduction phase, p. 287). The same is true
for the rare positional nystagmus and vertigo secondary to macroglobulinaemia.

Positional alcohol vertigo/nystagmus

(PAN)

Baniny (1911) described the direction-changing


characteristics of positional alcohol nystagmus in
humans with changes in head positions (beating
toward the undermost ear). This was later proven in
animals (Rothfeld 1913; De Kleyn and Versteegh
1930; Goldberg and Strtebecker 1941). The revers al
of direction in PAN (beating toward the uppermost
ear) hours after alcohol intake was first observed by
Walter (1954) and later termed PAN 11 by Aschan et

Alcohol (PAN I)

I~

t
t

~)

A lcohol (PAN 11)


Glycerol
Heavy water
Macroglobu I inaernia

horiz. ENG

~
Fig. 17.1. Ingestion of water-soluble molecules with differing
specific gravities, such as alcohol, heavy water or glycerol, causes
a specific gravity differential between cupula and endolymph
(buoyancy hypothesis) with positional nystagmus and vertigo.
During the resorption phase of alcohol, nystagmus beats toward
the undermost ear (PAN I with the cupula relatively lighter than
endolymph) . Positional nystagmus beats toward the uppermost
ear during alcohol-reduction phase (PAN 11) as weil as in glycerol-,
heavy water- and macroglobulinaemia-induced positional nystagmus (with the cupula relatively heavier than endolymph). The
gravity-dependent deflection force on the cupula (inset, B) must
be greater than the physiological restoring force (inset, C) for the
positional nystagmus to last as long as the precipitating head
position is maintained. (From Brandt 1990.)

al. (1956), Aschan (1958) and Money et al. (1965,


1974) in their studies. A peripheral labyrinthine
origin of PAN was suggested by observations that it
does not occur when labyrinthine function is lost in
humans (Harris et al. 1962) and animals (Nito et al.
1964).
Alcohol is lighter than endolymph, and when
blood levels approach 40 mg/dl - 1, alcohol diffuses
into the cupula, rendering it lighter than endolymph
and thereby transforming the semicircular canals
into gravity-sensitive receptors (Money et al. 1974).
Nystagmus and vertigo then occur when the subject
lies down. In phase I of PAN, the nystagmus beats
toward the undermost ear (Fig. 17.1). With time,
blood alcohol diffuses into the endolymph, equalis-

Positional nystagmus/vertigo with specific gravity differential between cu pu la and endolymph

ing its specifie gravity to that of the eupula. There is


then a "silent (intermediate) period", whieh beg ins
between 3.5 and 5 hours after eessation of alcohol
ingestion, when positional vertigo is absent. Alcohol
seleetively diffuses out of the eupula before it leaves
the endolymph. This eauses the eupula to be transiently denser than the endolymph, thus initiating
phase II of PAN, whieh begins between 5 and 10
ho urs after eessation of drinking, when blood levels
fall to about 20 mg/dl. In PAN II, nystagmus beats to
the uppermost ear. Positional vertigo may persist
until all alcohol eventually leaves the endolymph
(equalising the speeifie gravities of the endolymph
and eupula), but this may oeeur only many hours
after the blood alcohollevel has reaehed zero. PAN II
is usually assoeiated with "motion siekness" and is a
major eoneomitant of the hangover. The "morning
after" drink of alcohol may indeed re-equalise the
speeifie gravities and lessen, albeit transiently, the
untoward symptoms (Brandt and Daroff 1980).
The differential effeets of alcohol and "heavy
water" on vestibular funetion ean be easily investigated in animal experiments. It was shown in rabbits
that ethyl alcohol during PAN I deereased the
vestibular oeular reflex (VOR) gain and redueed the
long-time eonstant. Conversely, during PAN II ethyl
alcohol and "heavy water" enhaneed the VOR gain
and prolonged the long-time eonstant (Koizuka et al.
1989). These findings indieate that alcohol and
"heavy water" direetly alter the dynamics of the
eupula-endolymph system, making the semicireular
eanal the most probable site of the meehanism for
alcohol or "heavy water" -indueed positional nystagmus. However, the eentral effeets of alcohol on the
eentral nervous system, e.g. eerebellar struetures, are
also well known. Using 3-D eye movement analysis,
Fetter et al. (1999) showed that in addition to the
nystagmus indueed by the buoyaney of all six eupulae, alcohol intoxieation also eauses a vertieal offset,
whieh is independent of the orientation of the subjeet in spaee. This offset may have a toxie effeet on
eentral vestibular pathways and produee a tone
imbalanee of the VOR.
Experiments on the relationship between PAN
and postural imbalanee in healthy human volunteers
showed that the threshold for inereased body sway
was higher than the threshold for PAN I (Ledin and
dkvist 1991), although the intensity of PAN and
body sway showed a signifieant positive eorrelation
(Kubo et al. 1990). Posturographic measures of alcohol effeets are most sensitive when balancing with
the eyes closed. In an investigation of four age
groups of healthy men (20-59 years), PAN was not
signifieantly different (Jones and Neri 1994).

287

Table 17.1. Positional alcohol nystagmus (PAN)

Nystagmus

PANI

Time

Mechanism

Direction-changing
rotational vertigo
and nystagmus
with
head right or left,
beating toward
the undermost ear

30 min after oral With blood


levels 25-40
administration:
duration 3-4 h mg/dl
diffusion into
the cu pu la
makes it
lighter than
endolymph,
therefore
sensitive to
gravity changes

"Silent
intermediate
period"

No positional
vertigo nystagmus

3-5 h after
alcohol
ingestion

PAN 11

"Hangover vertigo" 5-10 h after


alcohol with
ingestion
direction-chang ing
positional nystagmus
with head right or
left, toward
the uppermost ear

(resorption
phase)

(reduction
phase)

Alcohol diffuses
also into the
endolymph:
equal specific
gravity of
cupula and
endolymph
Alcohol stays
longer in the
endolymph,
wh ich causes a
specific gravity
differential
with the cupula
being "heavier"

Positional"heavy water" nystagmus


Money and Myles (1974) reported that ingestion of
100-200 g deuterium oxide (DP) eaused a vigorous
lateral positional nystagmus lasting some ho urs in
humans, with a direetional eharaeteristie opposite to
that of postural alcohol nystagmus. Deuterium oxide
("heavy water") has a moleeular weight of 20.030
(water is 18.016) and a specifie weight of 1.10 (H 2 0 =
1.00).1t is thought to diffuse earlier into the eupula
than the endolymph. As long as a suffieiently large
specifie gravity differential is maintained between
the two, the eupula may aet inappropriately as a
gravity transdueer. In animal experiments "heavy
water" with a speeifie gravity of 1.10 also eaused an
apogeotropie type of positional nystagmus in rabbits
(Koizuka et al. 1989).

Positional glycerol nystagmus


Standard doses of glycerol, used to perform diagnostie audiograms in patients with suspeeted Meniere's
disease, ean eause transient positional nystagmus.

Vertigo

288

This was first observed in a single case by Angelborg


et al. (1971) and later studied more systematically by
Rietz et al. (1987): five of six subjects exhibited a
positional nystagmus shortly after peak serum levels
were achieved; a maximum was reached 120 minutes
after glycerol ingestion. From their data, Rietz et al.
(1987) inferred that different transport velocities of
the compound, first to the cupula, and then to the
endolymph, with a transient increase in density of
the cupula, result in a postural nystagmus toward the
uppermost ear. This concept is supported by animal
studies in guinea-pigs, in which Yoshida et al. (1985)
found that endolymphatic pressure dropped 15
minutes after intravenous administration of glycerol
(when glycerol enters the cupula?), lasting for at least
another 25 minutes, at which point glycerol enters
the endolymph and (re)balances the osmotic gradient. The duration of postural glycerol nystagmus,
most pronounced between 90 and 240 minutes after
oral ingestion (Rietz et al. 1987), indicates that
glycerol does not enter the endolymph for at least 2
hours after oral administration. Furthermore,
administration of glycerol can cause a small pressure
difference between the perilymph and the
endolymph (Takeda et al. 1990), dehydrate endolymphatic hydrops (Ho et al. 1993), and also affect
neuroactivity in isolated labyrinth preparations
(frog; increased resting discharge and decreased
evoked responses) (Zucca et al. 1995).

Positional nystagmus with


macroglobulinaemia (Waldenstrm's
disease)
Malignant lymphoproliferative Waldenstrm's disease with macroglobulinaemia can have auditory
and vestibular manifestations in as many as
10%-20% of patients (Logothetis et al. 1960; Fahey et
al. 1965). Various explanations have been proposed:
increased blood viscosity (Bolch and Maki 1973)
with obstruction in the venules (Andrews et al.
1988), stagnation or sudden release of dotting factors (Ronis et al. 1966; Ruben et al. 1969), associated
neurological dysfunction (Bolch and Maki 1973), or
haemorrhage (Afifi and Tawfeek 1971). Vascular vertigo in the hyperviscosity syndrome (see Chap. 21,
p. 341) is most likely to be due to obstruction of the
venules and capillaries with peripheral vestibular
hypoxia (Andrews et al. 1988). This mechanism does
not explain positional vertigo in Waldenstrm's
disease.
Keim and Sachs (1975) reported on five patients,

three of whom gave a history of periodic dizziness


with postural changes. They were able to record a
direction-changing positional nystagmus with a
latent period and fatigue in one patient, who simultaneously reported rotatory vertigo with nausea and
visual disturbance. The two others, symptom-free at
the time of otoneurological investigation, did not
exhibit positional nystagmus. Since gammaglobulins
with molecular weights exceeding one million are
possible in macroglobulinaemia, whereas normal
gammaglobulins have molecular weights of
approximately 150 000, Keim and Sachs (1975) stress
that the increased specific gravity of the cupula is
the causative factor, which varies with the concentration of the circulating protein and the diffusion
characteristics.

References
Afifi AM, Tawfeek S (1971) Deafness due to Waldenstrms
macroglobulinemia.) Laryngol Oto185:275
Andrews )C, Hoover LA, Lee RS, Honrubia V (1988) Vertigo in the
hyperviscosity syndrome. Otolaryngol Head Neck Surg
98:144-149
Angelborg C, Klockhoff I, Stahle) (1971) The caloric response in
Meniere's disease during spontaneous and glycerol-induced
changes of the hearing loss. Acta Otolaryngol (Stockh)
71:462-468
Aschan G (1958) Different types of alcohol nystagmus. Acta
Otolaryngol (Stockh) 140:69-78
Aschan G, Bergstedt M, Goldberg L, Laurell L (1956) Positional
nystagmus in man during and after alcohol intoxication. ) Stud
AlcohoI17:381-405
Barany R (1911) Experimentelle Alkoholintoxikation. Monatsschr
Ohrenheilkd 45:959-962
Bolch KJ, Maki DG (1973) Hyperviscosity syndrome associated
with immunoglobulin abnormalities. Semin Hematol
10:113-124
Brandt Th (1990) Positional and positioning vertigo and nystagmus.) Neurol Sci 95:3-28
Brandt Th, Daroff RB (1980) The multisensory physiological and
pathological vertigo syndromes.Ann NeuroI7:195-203
DeKleyn A, Versteegh C (1930) Untersuchungen ber den
sogenannten
Lagenystagmus
whrend
akuter
Alkoholvergiftung beim Kaninchen. Acta Otolaryngol (Stockh)
14:356-377
Fahey )L, Barth WF, Solomon A (1965) Serum hyperviscosity syndrome. )AMA 192:464-467
Fetter M, Haslwanter T, Bork M, Dichgans ) (1999) New insights
into positional alcohol nystagmus using 3-D eye movement
analysis. Ann Neurol In press
Goldberg L, Strtebecker TP (1941) Criteria of alcohol intoxication in animals in relation to blood alcohol. Acta Physiol Scand
3:71-81
Harris CS, Guedry FE, Graybiel A (1962) Positional alcohol
nystagmus in relation to labyrinthine function. NSAM 839,
NASA R-47, Naval School of Aviation Medicine, Pensacola
Ito M, Watanabe Y, Shojaku H, Kobayashi H, Aso S, Mizukoshi K
(1993) Furosemide VOR test for the detection of endolymphatic
hydrops. Acta Otolaryngol (Stockh) SuppI504:55-57
Iones AW, Neri A (1994) Age-related differences in the effects of

Positional nystagmus/vertigo with specific gravity differential between cupula and endolymph
ethanol on performance and behavior in healthy men. Alcohol
Alcoholism 29:171-179
Keim RJ, Sachs GB (1975) Position al nystagmus in association
with macroglobulinemia. Ann OtoI84:223-227
Koizuka I, Takeda N, Kubo T, Matsunaga T, Cha CI (1989) Effects
of ethyl-alcohol and heavy water administration on vestibuloocular reflex in rabbits. ORL 51:151-155
Kubo T, Sakata Y, Koshimune A, Sakai S, Ameno K, Ijiri I (1990)
Positional nystagmus and body sway after alcohol ingestion.
Am J Otolaryngol 11: 416 -419
Ledin T, dkvist LM (1991) Effect of alcohol measured by dynamic
posturography. Acta Otolaryngol (Stockh) SuppI481:576-581
Logothetis J, Silverstein P, Coe J (1960) Neurological aspects of
Waldenstrm's macroglobulinemia. Arch NeuroI3:564-573
Money KE, Myles WS (1974) Heavy water nystagmus and effects of
alcohol. Nature 247:404-405
Money KE, Johnson WH, Cerlett BMA (1965) Role of semicircular
canals in positional alcohol nystagmus. Am J Physiol
208: 1065-1070
Money KE, Myles WS, Hoffert BM (1974) The mechanism of positional alcohol nystagmus. Can J Otolaryngol 3:302-3l3
Nito Y, Johnson WH, Money KE, Ireland PE (1964) The nonauditory
labyrinth and positional alcohol nystagmus. Acta Otolaryngol
(Stockh) 58:65
Rietz R, Troia BW, Yonkers AJ, Norris TW (1987) Glycerol-induced

289

positional nystagmus in human beings. Otolaryngol Head Neck


Surg 97:282-287
Ronis ML, Rojer CL, Ronis BJ (1966) Otologic manifestations of
Waldenstrm's macroglobulinemia. Laryngoscope 76:5l3-523
Rothfeld J (1913) ber den Einflu akuter und chronischer
Alkoholvergiftung auf die Funktion des Vestibularapparates.
Monatsschr Ohrenheilkd 47: l392-l393
Ruben RJ, Distenfeld A, Berg P, Carr R (1969) Sudden sequential
deafness as the presenting symptom of macroglobulinemia.
JAMA 209: l364-1365
Schuknecht H (1969) Cupulolithiasis. Arch Otolaryngol
90:765-778
Takeda T, Takeuchi S, Saito H (1990) Effect of glycerol on pressure
difference between perilymph and endolymph. Acta
Otolaryngol (Stockh) 110:68-72
Takumida M, Hirakawa K, Harada Y (1995) Effect of glycerol on
the guinea pig inner ear after removal of the endolymphatic
sac. ORL J Otorhinolaryngol Relat Spec 57:5-9
Walter HW (1954) Alkoholmibrauch und Alkoholnystagmus.
Dtsch Z Ges Gericht! Med 43:232-241
Yoshida M, Lowry LD, Liu JJC (1985) Effects ofhyperosmotic solutions on endolymphatic pressure. Am J OtolaryngoI6:297-301
Zucca G, Maracci A, Milesi V, Trimarchi M, Mira E, Manfrin M,
Quaglieri S, Valli P (1995) Osmolar changes and neural activity
in frog vestibular organs. Acta Otolaryngol (Stockh) 115:34-39

Central positional vertigo

When the head is brought into an off-vertical, lateral


or head-hanging position, there is a change in graviceptive (otolithic) input. This change is the precipitating factor for central positional vertigo. The most
likely explanation for this response is that a vestibular tone imbalance occurs which manifests as directional positional nystagmus and rotatoryllinear
vertigo. It is probably caused by disinhibition of the
ve tibular reflexes to perception, eye, head or body
position (Brandt 1990).
Central positional nystagmus and/or vertigo indicate posterior fossa disorders that involve the loop
between the vestibular nudei and midline archicerebellar structures of the vermis. Before treating a
patient with positional signs and symptoms, the
physician must quickly distinguish between peripheral and central vestibular dysfunction, because the
latter necessitates an intensive battery of apparative
diagnostics in order to further differentiate among
the multiple causes of central vestibular dysfunction.
The most frequent differential diagnosis involves
benign paroxysmal positioning vertigo and central
positional/positioning vertigo or nystagmus. Four
characteristic types of central vestibular positionall
positioning syndromes can be delineated, although
some overlap in the signs and symptoms and combinations are possible in the same patient:

positional downbeating nystagmus (with or


without associated vertigo)
central positional nystagmus
central paroxysmal positional/positioning vertigo
central positioning vomiting

Central positional/positioning syndromes are less


common than benign paroxysmal positioning ver tigo (Chap. 16), and in most cases it is easy to make a
distinction on the basis of the typical features of
latency or duration and direction of the nystagmus
elicited by the precipitating head positioning
manoeuvre, the reversal of direction of the nystagmus on uprighting the patient, and the fatiguability
with repetitive stimulation (p. 253). However, this

distinction may prove difficult in individual cases,


since benign paroxysmal positioning vertigo and
central paroxysmal positioning vertigo can share
many of these features (Bttner et al. 1998). The following dinical rules appear to still be valid for identifying a central positional/positioning vestibular
syndrome:

a prolonged positional nystagmus (slow-phase


velocity >5/second) without a sociated vertigo
positioning-induced vomiting after single head
movements without associated intense positioning nystagmus
positionallpositioning vertigo and nystagmus
with pure torsional, down beat or upbeat directions (pure horizontal direction of nystagmus is
typical for peripheral horizontal canal BPPV,
p.269)
positional/positioning nystagmus not corresponding to the stimulated head (semicircular
canal) plane, Le. torsional nystagmus after horizontal canal stimulation.

The earlier rules that positional nystagmus beating


toward the uppermost or beating longer than 1
minute prove central pathology are no longer valid
(p. 270). Both features occur, e.g. with rare cupulolithiasis of the horizontal canal (Steddin and
Brandt 1996), and both can also indicate mixed
peripherallcentral dysfunction as in positional alcohol nystagmus (p. 286), drug intoxication or cerebellar degeneration. Table 18.1 summarises relevant
dinical features for the differential diagnosis of
peripheral benign paroxysmal positioning vertigo
and central paroxysmal positioning vertigo.

Positional downbeating nystagmus


Positional downbeating nystagmus, with only slight
vertigo in the head-hanging position, is indicative of

291

Vertigo

292
Table 18.1. Clinical features of peripheral benign paroxysmal positioning vertigo (BPPV) and central paroxysmal positioning vertigo (CPPV)
Feature

BPPV

CPPV

Latency following
precipitating
positioning

1-15 s
(shorter in h-BPPV)

1-5 s

Vertigo

Typical

Typical

Duration of attack

5-40 s

5->60 s

Nystagmus direction

Torsional-vertical with
head tilt in the plane of
the posterior (p-BPPV)
or the anterior
(a-BPPV) canal,
horizontal with
head rotation in the
plane of the horizontal
(h-BPPV) canal

Course of vertigo and


nystagmus in the
attack

Crescendo-decrescendo Crescendo-decrescendo
(with typical
is possible
canalolithiasis)

Nausea and vomiting

Rare on single
precipitating
manoeuvres
(if so, then associated
with exceptional
positioning nystagmus
intensity); not
uncommon with
serial provocation

Frequent on single
precipitating
manoeuvres
(not necessarily
associated
with strong nystagmus
intensity)

Natural course of the


condition

Spontaneous recovery
within days to months
in 70-80%

Dependenton
aetiology, spontaneous
recovery within
weeks in most cases

Associated
neurological
signs and symptoms

None (in idiopathic


BPPV)

Frequent cerebellar
and oculomotor signs
such as ataxia, saccadic
pursuit, gaze evoked
nystagmus, down beat
nystagmus, impaired
fixation suppression

Brain imaging

Normal

Lesions dorsolateral to
the fourth ventricle
and/or the dorsal
vermis (tumour,
haemorrhages,
infarctions or MS
plaques)
No specific lesions
(cerebellar
degeneration,
paraneoplastic
syndromes,
encephalopathy,
intoxication,
'idiopathic")

No latency or

manoeuvre

(Ionger in h-BPPV and


in rare cupulolithiasis)

Pure vertical or
torsional, combined
rotatory-linear,
direction changing; not
attributable to
stimulation of a
single canal aligned
to the plane of
precipitating head tilt

a vestibulocerebellar nodulus lesion. It may be related to the downbeat nystagmus syndrome, which also
shows activation on head extension (Chap. 11).
Experimental extirpation of the nodulus in the cat
causes postural downbeat nystagmus (Fernandez et
al. 1960), a finding that has also been confirmed by
clinical experience (Harrison and Ozsahinoglu 1972;
Kattah et al. 1984). Physiologically, the nodulus may
have an inhibitory influence on the gain of the
vertical vestibulo-ocular reflex (Fernandez and
Fredrickson 1964). Lesional postural downbeat
nystagmus can be abolished by additional bilateral
labyrinthectomy (Allen and Fernandez 1960).
Positional downbeat nystagmus, with or without
slight positional vertigo, is a frequent, often the only,
clinical sign in neurological patients. It may spontaneously resolve or persist and can be caused by
multiple sclerosis, ischaemia, intoxication, craniocervical malformation or cerebellar degeneration.
However, sometimes there is no identifiable aetiology
in elderly patients, and brain imaging techniques do
not even reveal a vestibulocerebellar lesion.

Central positional nystagmus


Positional nystagmus without concomitant vertigo is
always central; however, the direction of the nystagmus varies. Otoneurological examination of a total of
10730 patients by the same physician found central
positional nystagmus in 124 patients (1 %) (Ihomsen
et al. 1978). Ihis nystagmus may beat diagonally or
toward the undermost or uppermost ear. Frequently it
is bilateral and changes direction when the head is
tilted to the right or left. Ihe frequency of central
positional nystagmus is usually low and constant, distinguishing it from benign paroxysmal positional
nystagmus. Differentiation between central positional
nystagmus and benign paroxysmal positional nystagmus is not only based on the direction of the nystagmus but frequently also on the absence of a latency
period after movement to the provoking position. In
addition, there is generally a lack of fatiguability and
habituation on repetitive stimulation.
Central positional nystagmus is indicative of a
posterior fossa lesion, which is probably located in
the caudal brainstem or the vestibulocerebellum. A
more precise localisation is not possible. CI and
MRI are often unable to determine the location of
the lesion. Ihe possible causes are similar to those of
positional downbeating nystagmus. It is our experience that central positional nystagmus frequently
occurs in the elderly (lacunar ischaemia?) and often
resolves spontaneously.

293

Central positional vertigo

Fig.18.1.

Severe central positional vertigo is usually induced by infratentoriallesions dorsolateral of the fourth ventricle (CT scans

feft and centre of patients with cerebellar haemorrhages) or the vestibulocerebellum (cystic vermis tumour, right). (From Brandt 1990.)

Central paroxysmal positional/


positioning vertigo and paroxysmal
positioning vomiting
There is a variety of central - mostly paroxysmal positional vertigo syndromes. Linear or rotatory
vertigo, nystagmus, postural imbalance, nausea, and
vomiting may be abrupt and more violent than in
peripheral positional vertigo. This severe form of
central positional vertigo, which initially immobilises the patient - depending on its aetiology - gradually improves within days to weeks in most patients.
The following unilateral or bilateral lesions have
been reported to occur with central paroxysmal
positional vertigo:

b
Fig.18.2. Central paroxysmal positioning vertigo with nausea
and nystagmus beating toward the uppermost ear in a 4S-yearold patient with infarction of the posterior inferior cerebellar
artery on the right. Transverse (a) and sagittal (b) MRI sections
show infarction of the dorsal vermis. The sagittal section is
3.2 mm paramedian to the midline. The T2 signal-intense lesion
comprises ventrocaudal parts of the vermis, in particular uvula,
tonsils, lateral parts of the nodulus, and medial aspects of the
hemisphere. Associated ocular motor abnormalities included
saccadic pursuit, horizontal and vertical gaze-evoked nystagmus,
and impaired fixation suppression.

dorsolateral to the fourth ventricle including the


caudal cerebellar peduncles (Fig. 18.1; Brandt
1990)
dorsal vermis (Fig. 18.2) (Gregorius et al. 1976;
Watson et al. 1981; Barber 1984; Sakata et al.
1991)
diffuse cerebellar pathology (Bttner et al. 1999)
cerebellar drug intoxication (Arbusow et al.
1998).

The lesions dorsolateral to the fourth ventricle are


usually caused by haemorrhages (Fig. 18.1), tumours
or MS plaques. The lesions of the dorsal vermis are
caused by cerebellar midline tumours (e.g. medulloblastoma in childhood), and infarctions (e.g. posterior
inferior cerebellar artery; Fig. 18.2). Studies have
repeatedly reported (Drachman et al. 1977; Watson
et al. 1981) that the initial CT failed to show evidence
of a tumour or defined posterior fossa lesion which
was found only months (Drachman et al. 1977) or
years (Watson et al. 1981) later. Clinically such a

294

Vertigo

Fig. 18.3. Central paroxysmal positioning vertigo in a 44-year-old patient with a left-sided infarction of the posterior inferior cerebellar artery. Paroxysmal vertigo and nystagmus were elicited by positioning the head to the left or by shaking the head for 3-4 s in
the horizontal plane. Three-dimensional eye movement recordings (vertical, horizontal, torsional) after shaking the head in the horizontal plane. Eye position a, eye velocity b. Head shaking starts at the stippled verticalline and lasts for 1-2 s (black horizontal bar). This
leads without latency to an intense, predominantly torsional nystagmus, which lasts about 25 s. The torsional velocity reaches values
up to 150"/s. Eye movement recordings were performed in the light. (Bttner et al. 1999.)

delay can prove critical in cases of diffuse cerebellar


pathology, such as paraneoplastic syndrome, vasculitides of the nervous system or post-radiation
encephalopathy. Amiodarone, for example, when
given for drug-refractory tachyarrhythmias, can
cause a neurotoxic cerebellar syndrome (Palakurthy
et al. 1987) with ataxia, tremor, downbeat nystagmus,
and prolonged central vestibular positioninginduced vomiting (Arbusow et al. 1998) without any
pathology being evident in MRI.
Animal experiments have provided evidence that
posterior cerebellar vermis lesions may cause positional vertigo and nystagmus mimicking benign
paroxysmal positional vertigo (Allen and Fernandez
1960; Fernandez et al. 1960). The clinical relevance of
this central "pseudo" BPPV has been repeatedly
stressed (Gregorius et al. 1976; Watson et al. 1981;
Sakata et al. 1991); however, it should not be overestimated, for it is very rare.
The small number of experimental animal studies
and the relatively few well-documented clinical case
reports allow only speculation on the structures and
mechanisms causing central paroxysmal positioning
vertigo. It is obviously related to archicerebellar disinhibition of positional vestibular reflexes and
otolith-semicircular interaction. The critical vestibular structures include the dorsal vermis with the
nodulus and a vestibular nuclei-archicerebellar loop
that passes through the cerebellar peduncles.
The following clinical features have been
described for central paroxysmal positioning vertigo,
which is sometimes difficult to separate from
canalolithiasis or cupulolithiasis of the semicircular
canals (Bttner et al. 1998):

Latency
There is virtually no latency (Gregorius et al.
1976; Jacobsen et al. 1995) or latencies up to 3-5 s
(Watson et al. 1981). In animal experiments with
lesions of the nodulus latencies varied between
0-50 s (Allen and Fernandez 1960; Fernandez et
al. 1960).
Duration of the attack
In patients an attack can last from a few seconds
(5-6 s, Barber 1984; 15 s, Gregorius et al. 1976) to
less than 1 minute (Drachman et al. 1977). In
experimental animaliesions of the nodulus a
duration of 30-180 s was found (Fernandez et al.
1960).

Fatiguability on repetitive positioning


This has been seen in patients (Drachman et al.
1977; Kattah et al. 1984; Gregorius et al. 1976)
and in animals with experimental dorsal vermis
lesions (Allen and Fernandez 1960). The absence
of fatiguability (Gregorius et al. 1976; Sakata et
al. 1991) has also been reported.
Crescendo-decrescendo course of the attack
This was reported in patients (Sakata et al. 1991)
and experimental animals (Fernandez et al.
1960).
Vomiting
The occurrence of vomiting on single positioning manoeuvres without intense nystagmus
(Drachman et al. 1977) reflects our own experience with central paroxysmal positional vertigo,
namely that there may be a dissociation between
nausea and vomiting, on the one hand, and vertigo and nystagmus, on the other. In peripheral
benign paroxysmal positioning vertigo vomiting
is usually correlated with nystagmus intensity
and is elicited only by repetitive provocation.
Natural course
Spontaneous recovery after weeks has been
reported in patients with positioning-evoked
vomiting (Drachman et al. 1977) and for dorsal
vermis lesions in experimental animals (Allen
and Fernandez 1960).
Associated clinical features
Central paroxysmal positional vertigo has been
described in patients with otherwise normal eye
movements (Barber 1984) and a completely normal neurological evaluation (Gregorius et al.
1976; Watson et al. 1981; Sakata et al. 1991).
However, most of the thus afflicted patients show
additional ocular motor abnormalities (saccadic
pursuit, gaze-evoked nystagmus, downbeat
nystagmus, impaired fixation suppression) or
cerebellar ataxia.

In the few instances where central paroxysmal positional vertigo may mimic peripheral benign
paroxysmal positional vertigo, the direction of
nystagmus after the positioning manoeuvre can provide clear signs of a central dis order (Fig. 18.3;
Bttner et al. 1999). The nystagmus resulting from
peripheral canalolithiasis or cupulolithiasis always
beats with a horizontal or vertical-torsional pattern,

Central positional vertigo

295

eye position (deg)


down

40

-------------,-------------------------------------------------------------------------------------------

30
vertical

20
10
up

left

o
-10
30

horizontal

20
10

right

CW

0
-1 0

30

torsiona l

20
10
0

CCW

-10
-20 ~-------4--L-----~1~0--------~15--------~2-0------~2~5------~30 time (s)

eye velocity (degfs)


down

500

300

T-- . --- . ---. ---'---,----------- ------------ ----------------------------------- - --------- --------- --- ------- ---

vertical

100
-100

up

-300

300

left

horizontal

100
-1 00

right

cw

-300

300

torsional

100
-1 00

ccw

- 300
-500 ~------~--~----4_--------~------~--------+_------~
o
5
10
5
20
25
30

time (s)

Vertigo

296

which reflects ampullofugal or ampullopetal stimulation of the affected horizontal or vertical semicircular canal. The plane of the affected canal must
be aligned with the plane of the provocative head
tilt.
The management of patients with acute severe
central paroxysmal positioning vertigo poses a considerable challenge. If vomiting occurs several times
a day after single head positionings, we recommend
the prescription of bedrest and avoidance of head
movements as much as possible. Antivertiginous
drugs, e.g. dimenhydrinate or scopolomine, are helpful to some extent, but they do not have the desired
preventative effect in all patients. Some of our
patients required intravenous nutrition for a few
weeks. In this unsatisfying state of affairs, it is
important to inform the patient about the natural
course of the vertigo, which tends to resolve within
days to weeks. Baclofen and clonazepam can be
tried, but further drug studies are needed to find
more effective remedies that reduce the overexcitability of vestibular positioning reflexes or at
least halt the vomiting. The obviously dissociated
efficacy of benzodiazepines on vomiting but not on
positioning nystagmus in some of our patients suggests that vomiting in these cases was elicited directly
by dorsal vermis projections to the reticular formation (vomiting centre, p. 485) rather than by dorsal
vermis projections to the vestibulo-ocular reflex
(ArbusQw et al. 1998).

Transient vertebrobasilar ischaemia


Nystagmus, vertigo, and postural imbalance are frequently reported. They are induced when the head is
maximally rotated and/or extended while standing,
and they terminate abruptly when the head is
returned to anormal upright position. However,
they may indicate neurovascular cross-compression
of the eighth nerve (p. 117) or transient vertebrobasilar ischaemia caused by a functional compression of the vertebral artery, particularly in
elderly patients with atheromas, cervical spondylosis
or osteophytes that narrow the transverse foramina
(Denny-Brown 1960; Williams and Wilson 1962;
Sheehan et al. 1960).
Episodic vertigo and ocular motor abnormalities
are common early symptoms of reduced vertebrobasilar blood flow due to the steep pressure
gradient from the aorta to the terminal pontine
arteries. These long, circumferential arteries provide
a highly vulnerable blood supply for the vestibular
nuclei (Williams and Wilson 1962). Experimental

studies on blood flow in cadavers have revealed that


extreme head positions may reduce flow through
one or another of the vertebral or carotid vessels
(Toole and Tucker 1960). Transient ischaemia of the
labyrinths mayaiso playa role, since the blood supply originates from the same source.
A partial obstruction of the arteries may combine
with a sudden fall of systemic blood pressure, e.g.
when a patient rises from achair and looks upward.
However, isolated episodes of vertigo without concurrent neurological signs and symptoms are an uncommon manifestation of vertebrobasilar ischaemia
(Fisher 1967; Estol et al. 1996). They should suggest
other disorders, such as benign paroxysmal positional
vertigo, vestibular paroxysmia or perilymph fistula.
An association of dysfunctions that confirms the
diagnosis is vertigo with visual illusions, field defects,
diplopia, dysphagia, dysarthria, drop attacks (see also
Vestibular drop attacks in Meniere's disease, p. 94) or
motor symptoms.

Rotational vertebral artery occlusion


If transient ischaemic attacks with vertigo, nystagmus, and ataxia occur secondary to vertebral artery
compression with rotational head motion, this is
termed the syndrome of rotational vertebral artery
occlusion (Yang et al. 1985; Kuether et al. 1997).
Vertebral arteries are susceptible to mechanical
compression with head rotation, because of muscular and tendinous insertions (Powers et al. 1961;
Mapstone and Spetzler 1982; Kojima et al. 1985) and
osteophytes and other degenerative changes of cervical spondylosis. When the artery passes through
the foramina transversaria of the cervical vertebrae
at C6-C2, the patients may be symptomatic on right
and/or left head rotation (Sheehan et al. 1960; Bakay
and Leslie 1965; Kuether et al. 1997). Many of them
also have one hypoplastic vertebral artery. Transient
occlusion of the artery bears the risk of formation of
a thrombus, which may embolise toward the vertebral basilar territory or even cause vertebral artery
or basilar thrombosis (Grossman and Davis 1982;
Okawara and Nibbelink 1974).
A review of the literature on rotational vertebral
artery occlusion reveals that vertigo, dizziness,
nystagmus and ataxia are the leading presenting
symptoms and that surgical treatment is effective
and should be considered to avoid further morbidity (Kuether et al. 1997). The correct site of occlusion
of the vertebral artery can be demonstrated only by
dynamic angiography during progressive stepwise
head rotation. We agree with Kuether and colleagues (1997) that the preferred occlusion of the

297

Central position al vertigo


eyes open
head upright
20 Nm
20 Nm

head extension

fore-aft

QI lateral

Q~--------------------~-----~--------------postural
sway

eyes closed

eyes open

~.,~~~
postural
sway
(foam rubber
platform)

Fig.18.4.

Physiological head extension vertigo. Differential effects of head extension and normal head position on fore-aft and lateral
body sway (original recordings), with the eyes open or closed. Normal subject standing on a firm posturography platform (top) or on a
slice of foam rubber (bottom). Postural imbalance is the most pronounced when the subject is standing on foam rubber during head
extension with the eyes closed. (From Brandt et al. 1981.)

vertebral artery occurs at the C2 level. Foraminal


decompression is possible either by a posterior lateral approach with partial transversectomy and
unroofing of the foramina or by a midline approach
with discectomy and osteophyte removal (Kuether
et al. 1997). The conservative approach consists of
anticoagulation and avoidance of symptomatic
head turning.

Head (neck)-extension vertigo


If vertebrobasilar ischaemia is excluded, symptoms
of to-and-fro vertigo and postural imbalance frequently occur in healthy persons. For example, when
they are doing overhead work and standing on
unstable, wobbly ladders or in situations where the

visual cues confiict with proprioceptive input (looking up at moving clouds). Vertigo and postural
imbalance terminate abruptly when the head is
fiexed to a neutral position. These symptoms are
also often attributed to intermittent vertebral artery
occlusions caused by the head posture, particularly
in elderly persons; however, they frequently also
occur in young people. A physiological explanation
is based on an unusual combination of multisensory
inputs from the stabilising systems (Brandt and
Daroff 1980; Brandt et al. 1981). The "normal" instability related to this head position (Fig. 18.4) can be
easily demonstrated by attempting to balance on one
foot with the eyes closed and head extended as compared with a neutral head position.
In such head-extended positions, the otoliths
must operate outside their optimal range (Brandt et
al. 1981). When the body is supine, the otoliths are in
the same position relative to the gravitational vector;

298

however, they are not then involved in postural control. Also during head (neck) flexion, the otoliths
must function outside their optimal range, but this
does not cause a considerable postural instability.
Flexion is a common position that is consequently
better adapted to. Repetitive challenges are required
for the development of central adjustments of motor
responses to the patterns of multisensory inputs.
Infrequently occurring stimulation patterns would
not be associated with appropriately calibrated sensorimotor engrams.
Visual cues, which correct for postural imbalance,
would be less effective if the head is extended,
because the egocentric spatial coordinates change
(with respect to retinal shift, up-down becomes foreaft). Moreover, the direction of compensatory body
sway must be corrected to reflect the change in coordinates. In situations in which visual cues are absent
(eyes closed or in darkness) or conflicting (looking
up at moving clouds), the postural imbalance is
greatly worsened. When somatosensory input varies,
for example, when a person is standing on a piece of
foam rubber or in patients with sensory polyneuropathy (Figs 30.2 and 30.3), the symptoms of
instability also increase.

Bending-over vertigo
A combination of mechanisms similar to that
described for head-extension vertigo would explain
the common symptom of vertigo on bending over at
the waist (Brandt and Daroff 1980). An added consideration is the transient increase in intracranial
pressure (secondary to increased cephalic venous
pressure) which is transmitted to the perilymphatic
space surrounding the endolymphatic membrane.

References
Allen G, Fernandez C (1960) Experimental observations in postural nystagmus: extensive lesions in posterior vermis of the
cerebellum. Acta Otolaryngol (Stockh) 51:2-14
Arbusow V, Strupp M, Brandt Th (1998) Amiodarone induced
severe prolonged head positional vertigo and vomiting.
Neurology 51:917
Bakay L, Leslie EV (1965) Surgical treatment of vertebral artery
insuffieiency caused by cervical spondylosis. J Neurosurg
23:596-602
Barber HO (1984) Positional nystagmus. Otolaryngol Head Neck
Surg 92:649-655
Brandt Th (1990) Positional and positioning vertigo and nystagmus. J Neurol Sei 95:3-28

Vertigo
Brandt Th, Daroff RB (1980) The multisensory physiological and
pathological vertigo syndromes. Ann NeuroI7:195-203
Brandt Th, Krafczyk S, Malsbenden I (1981) Postural imbalance
with head extension: Improvement by training as a model for
ataxia therapy.Ann NY Acad Sei 374:636-649
Bttner U, Heimchen Ch, Brandt Th (1998) Diagnostic criteria for
central versus peripheral positioning nystagmus. Acta
Otolaryngol (Stockh) (in press)
Bttner U, Brandt Th, Heimchen Ch (1999) The direction of nystagmus is important for the diagnosis of central paroxysmal
positioning nystagmus (cPPV). Neuroophthalmology (in
press)
Denny-Brown D (1960) Recurrent cerebrovascular episodes. Arch
NeuroI2:194-209
Drachman DA, Diamond ER, Hart CW (1977) Posturally evoked
vomiting: assoeiation with posterior fossa lesions. Ann Otol
Rhinol Laryngol 86:97-101
Estol C, Caplan LR, Pressin MS (1996) Isolated vertigo: An
uncommon manifestation of vertebrobasilar ischaemia.
Cerebrovasc Dis 6 (SuppI2): 161
Fernandez C, Fredrickson JM (1964) Experimental cerebellar
lesions and their effect on vestibular function. Acta Otolaryngol
(Stockh) 192:52-62
Fernandez C, Alzate R, Lindsay JR (1960) Experimental observation on postural nystagmus. Ir. Lesions of the nodulus. Ann
Otol Rhinol Laryngol (Stockh) 69:94-114
Fisher CM (1967) Vertigo in cerebrovascular diseases. Arch
Otolaryngol Head Neck Surg 85:529-534
Gregorius FK, Grandall PH, Baloh RW (1976) Positional vertigo
with cerebellar astrocytoma. Surg NeuroI6:283-286
Grossman RI, Davis KR (1982) Position al occlusion of the vertebral artery: a rare cause of embolic stroke. Neuroradiology
23:227-230
Harrison MS, Ozsahinoglu C (1972) Positional vertigo: aetiology
and clinical significance. Brain 95:369-372
Jacobsen GP, Butcher JA, Newman CW, MonseIl EM (1995) When
paroxysmal positioning vertigo isn't benign. J Am Acad Audiol
6:346-349
Kattah JC, Kolsky MP, Luessenhop AJ (1984) Positional vertigo
and the cerebellar vermis. Neurology 34:527-529
Kojima N, Tamaki N, Fujita K, Matsumoto S (1985) Vertebral
artery occlusion at the narrowed "scalenovertebral angle":
mechanical vertebral occlusion in the distal first position.
Neurosurgery 16:672-674
Kuether TA, Nesbit GM, Clark WM, Barnwell SL (1997) Rotational
vertebral artery occlusion: a mechanism of vertebrobasilar
insuffieiency. Neurosurgery 41:427-433
Mapstone T, Spetzler RF (1982) Vertebrobasilar insuffieiency secondary to vertebral artery occlusion from a fibrous band. J
Neurosurg 56:581-583
Okawara S, Nibbelink D (1974) Vertebral artery occlusion following hyperextension and rotation of the head. Stroke 5:640-642
Palakurthy PR, Iyer V, Meckler RJ (1987) Unusual neurotoxieity
associated with amiodarone therapy. Arch Intern Med
147:881-884
Powers SR, Drislane TM, Nevin S (1961) Intermittent vertebral
artery compression: a new syndrome. Surgery 49:257-264
Sakata E, Ohtsu K, Itoh Y (1991) Positional nystagmus of benign
paroxysmal type (BPPN) due to cerebellar vermis lesions.
Pseudo-BPPN.Acta Otolaryngol (Stockh) SuppI481:254-257
Sheehan S, Bauer RB, Meyer JS (1960) Vertebral artery compression in cervical spondylosis. Neurology 10:968-986
Steddin S, Brandt T (1996) Horizontal canal benign paroxysmal
positioning vertigo (h-BPPV): transition of canalolithiasis to
cupulolithiasis. Alm NeuroI40:918-922
Thomsen J, Zilstorff K, Johnsen NJ (1978) Positional nystagmus of
the persistent type. ORL J Otorhinolaryngol Relat Spec
40:86-91

Central positional vertigo


Toole JF, Tucker SH (1960) Influence ofhead position upon cerebral circulation. Arch NeuroI2:616-623
Watson CP, Terbrugge K (1982) Positional nystagmus of the
benign paroxysmal type with posterior fossa medulloblastoma.
Arch NeuroI39:601-602
Watson CP, Barber HO, Peck J, Terbrugge K (1981) Position al vertigo and nystagmus of central origin. Can J Neurol Sei
8:133-137

299
Williams D, Wilson TG (1962) The diagnosis of the major and
minor syndromes ofbasilar insuffieiency. Brain 85:741-777
Yang PI, Latack JT, Gabrielsen TO, Knake JE, Gebarsk SS, Chandler
WF (1985) Rotational vertebral artery occlusion at C l-C2. AJNR
Am J NeuroradioI6:98-100

SECTION E
Vascular vertigo

Ischaemia can be responsible for a wide range of


vestibular syndromes. Most of the central and some
of the peripheral vestibular syndromes listed in
Table E.1 may have a vascular aetiology.
Nevertheless, most of these syndromes are covered
elsewhere, because there are many other possible
aetiologies, and a correct differential diagnosis is
very important.
Ischaemia will sometimes produce a combination
of peripheral and central symptoms, for example in
PICA (posterior inferior cerebellar artery) and AICA
(anterior inferior cerebellar artery) infarctions
(p. 308). The PICA and the AICA (Table E.2) supply
overlapping territories in the brainstem and cerebellum, and the AICA also supplies the peripheral
labyrinth via the internal auditory end artery. The
anterior vestibular branch of the internal auditory
artery, which supplies the superior part of the
vestibular labyrinth (including the horizontal and
anterior semicircular canal, but sparing the posterior
semicircular canal and utricle), seems to be particularly vulnerable to ischaemia (p. 308). A syndrome
may result which cannot be distinguished from
vestibular neuritis. A similar syndrome may also be
induced by partial PICA or AICA infarctions presenting with associated cerebellar signs (p. 309).
In migraine (p. 329) and in vertebrobasilar artery
ischaemia ("vertebrobasilar insufficiency", Williams
and Wilson 1962) it may not be possible to determine whether the vertigo is of peripheral or central
origin. While the course and prognosis of vascular
vertigo vary, the symptoms are transient when due
to ischaemic attacks. Thus, these symptoms have to
be differentiated from conditions exhibiting episodic
vertigo such as Meniere's disease, migraine with vertigo aura, vestibular paroxysmia (p. 117),vestibular
epilepsy, and others. As Fisher (1967) already
emphasised, isolated episodes of vertigo without
other neurological symptoms are an uncommon
manifestation of vertebrobasilar ischaemia and
should suggest other dis orders (Estol et al. 1996), e.g.
basilar migraine (Dieterich and Brandt 1999). The

Table E.l. Mechanisms and sites ofvascular vertigo


Syndrome

Site

Migraine

Basilar migraine
Benign paroxsymal
vertigo of childhood
Benign recurrent
vertigo

Pontomedullary
brainstem,
vestibulocerebellum
and/or
labyrinth

Infarction or
haemorrhage

Vestibular or hearing
loss (infarction of
AICA or internal
auditoryartery)
Pseudo "vestibular
neuritis" (with lacunar
or PICA infarction)
Ocular tilt reaction
(pontomesencephalic
or medullary infarction,
Wallenberg's syndrome)
Lateropulsion
(Wallenberg's syndrome)
Downbeat nystagmus/
vertigo

Labyrinth, vestibular
nerve

Mechanism
---~~~

-----

Upbeat nystagmus/
vertigo
Thalamic astasia
Central positional
vertigo
Central postional
nystagmus
Positional down beat
nystagmus

Vestibular nerve root or


vestibular nuclei
Pontomesencephalic
(contraversive);
pontomedullary
(ipsiversive)
Dorsolateral medulla
Vestibular nuclei commissure or
flocculus
Pontomesencephalic
junction or caudal
medulla
Posterolateral thalamus
Vestibular nuclei vestibulocerebellar
loop
?
Nodulus

Neurovascular cross- Vestibular paroxysmia


("disabling positional
compression
vertigo")

Vestibular nerve

Vascular polyneuropathy

Diabetic vestibular loss

Vestibular nerve,
labyrinth?

Hyperviscosity with
venous obstruction

Hyperviscosity syndrome Labyrinth


with episodic vertigo

Miscellaneous: with
secondary signs

Decompression sickness, Labyrinth


labyrinthine haemorrhage,
perilymph fistula or
hydrops, vascular
fistula signs

303

Vertigo

304
Table E.2. Mechanism of symptoms and signs commonly seen with
infarction in the distribution areas of the PICA and the AICA
Symptoms/signs

Structures involved
with PICA infarct

Structures involved
with AICA infarct

Vertigo, nystagmus

Vestibular nuclei,
posteroinferior
cerebellum

Labyrinth, vestibular
nerve, flocculus

Tinnitus, hearing
loss

None

Cochlea, auditory nerve,


cochlear nucleus

Gait and limb ataxia

Ventral spinocerebellar
tract, posteroinferior
cerebellum

Middle cerebellar
peduncle, anterior
inferior cerebellum

Dysphagia,
decreased gagging

Vagal nuclei and nerve

None

Facial
hemianaesthesia

Fifth cranial nerve and


nucleus

Cranial nerve V and


nucleus

Facial para lysis

Seventh cranial nerve

Cranial nerve VII

Crossed hemisensory Spinothalamic tract


loss
Horner syndrome

Spinothalamic tract

Descending sympathetic Descending sympathetic


fibres
fibres

PICA, posteroinferior cerebellar artery; AICA, anteroinferior cerebellar


artery.
From Baloh (1996).

most frequent symptoms associated with vertigo in


vertebrobasilar ischaemia are visual ones (diplopia,
visual illusions and pseudohallucinations, field
defects), but also drop attacks, unsteadinessl incoordination, extremity weakness, confusion, headache,
hearing loss, and dysarthria occur (Grad and Baloh
1989).

The extreme manifestations of central vascular


vertigo are either sudden, incapacitating, severe
rotational vertigo or severe positional vertigo
(p. 291). A vascular pathology is often considered
only in patients who present with positional nystagmus/vertigo (p. 292) but whose CT or MRI examinations do not indicate a corresponding lesion.
Nystagmus, vertigo, and postural imbalance, which
are induced when the head is maximally rotated
and/or extended while standing, and terminated
abruptly by returning the head to a neutral upright
position, are frequently reported. Clinicians usually
attribute these symptoms to inter mitte nt vertebrobasilar ischaemia caused by a functional compression of the vertebral artery. Particularly in
elderly patients and patients with unilateral
hypoplasia of one vertebral artery, this can be due to
atheromas, cervical spondylosis or osteophytes,
which narrow the transverse foramina (DennyBrown 1960; Williams and Wilson 1962; Sheehan et
al. 1960).

Episodic vertigo is the most common early symptom of reduced vertebrobasilar blood flow due to the
steep pressure gradient from the aorta to the terminal
pontine arte ries. These long, tenuous, circumferential arte ries provide a highly vulnerable blood supply for the vestibular nuclei (Williams and Wilson
1962). The possibility cannot be excluded that transient ischaemia of the labyrinth mayaiso playa role,
since its blood supply originates from the same
source. Experimental studies on blood flow in cadavers
revealed that extreme head positions may reduce
flow through one or the other of the vertebral or
carotid vessels (Toole and Tucker 1960).
Certain types of infarctions cause specific syndromes, for example, ipsiversive lateropulsion and
ocular tilt reaction in the case of dorsolateral
medullary infarction (Wallenberg's syndrome), or
contraversive ocular tilt reaction in paramedian
thalamomesencephalic infarction. However, haemorrhages, inflammation or acute space-occupying
lesions are other principal causes of the same
syndromes.
Intracerebral haemorrhages cause central vestibular syndromes usually at the three following locations: in the thalamomesencephalic region
(contraversive ocular tilt reaction), the pons, and the
paramedian vestibulocerebellar structures dorsolateral of the fourth ventricle (positional vertigo).
Cortical infarctions within the middle cerebral
artery territory may result in spatial disorientation,
lateropulsion (p. 192), and, only in exceptional cases,
in transient vertigo (Brandt et al. 1995). Despite
vestibular projections to the parietotemporal cortex,
these vestibulothalamocortical projections do not
carry tonic signals but rather integrate vestibular,
proprioceptive, and visual signals, providing a conscious awareness of body orientation (Baloh 1992).
Therefore, vertigo in cerebrovascular disorders is a
reliable localising symptom for critical vertebrobasilar
rather than carotid circulation.
Presyncopallight-headedness resulting from diffuse cerebral ischaemia in orthostatic hypotension,
cardiac arrhythmia or hyperventilation is nonlocalising and not a symptom of impending stroke
(Baloh 1992).
Two interesting vascular pathomechanisms
should be mentioned here, for which it is unfortunately not possible to definitively confirm a diagnosis:
neurovascular cross-compression with head motion
intolerance (vestibular paroxysmia, Brandt and
Dieterich 1994; "disabling positional vertigo", M011er
et al. 1986; p. 117), and hyperviscosity syndrome
(Andrews et al. 1988) with venous obstruction of
labyrinthine blood flow (p. 341). Appreciation of the
true incidence of sinus and cerebral venous thrombosis (pathologies ignored for many years) has led

Vascular vertigo
to the recognition that venous obstruction is the second most important cause of ischaemia (after arterial
occlusion), not only in the cerebrum and spinal cord,
but probably also in the labyrinth. The rare condition of episodic peripheral vestibulopathy in sickle
ceH disease (Savundra et al. 1996) can serve as an
example; it can be caused by either arte rial or
venous occlusion due to aggregated sickled erythrocytes.
FinaHy, decompression sickness (p. 352) can also
be considered a kind of vascular vertigo syndrome.
Labyrinthine haemorrhage can cause perilymph fistulas (p. 99) or delayed endolymphatic hydrops
(p. 88). Vascular signs of provoked vertigo may help
to establish the diagnosis of perilymph fistulas
(p. 99).

References
Andrews J, Hoover LA, Lee RS, Honrubia V (1988) Vertigo in the
hyperviscosity syndrome. Otolaryngol Head Neck Surg
98:144-149

305
Baloh RW (1992) Stroke and vertigo. Cerebrovasc Dis 2:3-10
Baloh RW (1996) Vestibular dis orders due to cerebrovascular disease. In: Baloh RW, Halmagyi GM (eds) Disorders of the
vestibular system. Oxford University Press, Oxford, pp 418-429
Brandt Th, Dieterich M (1994) Vestibular paroxysmia: vascular
compression of the eighth nerve. Lancet 343:798-799
Brandt Th, Btzel K, Yousry T, Dieterich M, Schulze S (1995)
Rotational vertigo in embolie stroke of the vestibular and auditory cortices. Neurology 45:42-44
Denny-Brown D (1960) Recurrent cerebrovascular episodes. Arch
NeuroI2:194-209
Dieterich M, Brandt Th (1999) Episodic vertigo related to
migraine (90 cases): vestibular migraine? J Neurol (submitted)
Estol C, Caplan LR, Pressin MS (1996) Isolated vertigo: An
uncommon manifestation of vertebrobasilar ischaemia.
Cerebrovasc Dis 6 (SuppI2): 161
Fisher CM (1967) Vertigo in cerebrovascular diseases. Arch
Otolaryngol Head Neck Surg 85:529-534
Grad A, Baloh RW (1989) Vertigo of vascular origin: Clinical and
ENG features in 84 cases.Arch NeuroI46:281-284
Melller MB, Melller AR, Jannetta PI, Sekhar L (1986) Diagnosis and
surgical treatment of disabling positional vertigo. J Neurosurg
64:21-28
Savundra P, Skacel P, Rudge P (1996) Peripheral vestibulopathy in
sickle cell disease. J Audiol Med 5:61-66
Sheehan S, Bauer RB, Meyer JS (1960) Vertebral artery compression in cervical spondylosis. Neurology 10:968-986
Toole JF, Tucker SH (1960) Influence ofhead position upon cerebral circulation. Arch NeuroI2:616-623
Williams D, Wilson TG (1962) The diagnosis of the major and
minor syndromes ofbasilar insufficiency. Brain 85:741-777

Stroke and vertigo

Strokes causing peripheral and


central vestibular disorders

Table 19.1. Classification of some central vestibular syndromes of the

brainstem tegmentum according to the three major planes of action of the


VOR: yaw, pitch, and roll' (for further description, see p. 167)

....:...._----

The discussion of vascular vestibular syndromes in


this chapter is restricted to clearly defined strokes of
the labyrinth, the vertebrobasilar, and the middle
cerebral artery territories. These vascular territories
supply the peripheral and central vestibular pathways and integration centres (Dieterich and Brandt
1995; Baloh 1996).
Vestibular pathways run from the eighth nerve
and the vestibular nuclei along the medial longitudinal
fascicle to the oculomotor nuclei and the supranuclear integration centres in the rostral midbrain.
From there, they reach several vestibular cortex
areas through thalamic projections. Most central
ischaemic vertigo syndromes are secondary to paramedian infratentorial lesions. Supratentorial
vestibular syndromes are less frequent, and only
those of the thalamus and the vestibular cortex are
relevant to the present discussion.
In the following, an attempt is made to correlate
circumscribed unilateral infarctions of the brainstern, the thalamus, and the temporoparietal cortex
with the clinical syndromes and the particular
vestibular structures involved. Such a scheme of
extended vestibular pathways, with different lesion
sites causing characteristic vestibular syndromes, is
required for topographic diagnosis (see Section C,
p.178).

Each pathway that mediates the vestibulo-ocular


reflex (VOR) in one of the three major planes of
action must run a separate course in order to
become individually lesioned (Brandt 1991; Leigh
and Brandt 1993). A classification has been proposed
for central vestibular disorders of the brainstem
(Table 19.1), wh ich attributes vestibular syndromes
such as downbeat or upbeat nystagmus,
pseudovestibular neuritis, and ocular tilt reaction
(OTR) to lesional tone imbalances in one of the three
major planes of action of the VOR (Brandt 1991;

Disorders of the VOR in the


horizontal
(yawl plane
Disorders of the VOR in the
sagittal (pitchl plane
Disorders of the VOR in the
frontal (roll) plane

Horizontal nystagmus:
pseudovestibular
neuritis (partial AICA/PICA infarctions;
multiple sclerosis plaques)
Vertical nystagmus: downbeat
nystagmus, upbeat nystagmus
Ocular tilt reaction and its components

aSyndromes in roll can indude complete OTR or its single components,


such as skew deviation, ocular torsion, lateral falls, and tilt of perceived
vertical. Syndromes in yaw indude spontaneous horizontal nystagmus,
rotation al vertigo, postural imbalance, and lateral falls. Syndromes in pitch
indude upbeat/downbeat nystagmus, pitch tilt of subjective vertical, and
fore-aft postural instability.

Brandt and Dieterich 1995) (See p. 169 and Chaps.


10-12). This hypothetical classification is based on
clinical experience and on animal studies showing
that a given syndrome can be elicited by distinct and
separate lesions of the pathways at different brainstern levels. In fact, pathways that mediate the VOR
in one of the three major planes run separate courses
and can becorne lesioned individually or in combination. Our approach to vestibular syndromes in
stroke (Dieterich and Brandt 1995) is based on
defined vascular territories and the evaluation of
lesional malfunction relative to the

307

perceptional,
ocular motor, and
postural effects in the
- yaw,
- pitch, and
- roll planes.

Vertigo

308

Anterior inferior cerebellar artery and


the internal auditory artery
The anterior inferior cerebellar artery (AICA) not
only supplies the anterolateral pons, the middle
cerebellar peduncle, and the flocculus, but also the
inner ear (Atkinson 1949; Kim et al. 1990). Thus,
ischaemia of the peripherallabyrinthine receptors,
the eighth nerve, the vestibular nucleus, or the
vestibulocerebellum could cause vertigo. The clinical
spectrum includes brainstem and cerebellar signs
and symptoms (Amareneo 1991), but labyrinthine
infarction has also been demonstrated histopathologically (Hinojosa and Kohut 1990).
Oas and Baloh (1992) described two patients with

clinical features of AICA infarctions who had had


recurrent attacks of rotatory vertigo. The vertigo initially appeared as an isolated symptom months prior
to the brain infarction and recurred at the time of
the infarction, accompanied by unilateral hearing
loss, tinnitus, facial numbness and hemiataxia.
Unilateral auditory and vestibular dysfunctions on
the affected side were documented by audiometry
and caloric irrigation. Thus, the authors reasoned
that the vertigo attacks preceding the infarction
resulted from transient ischaemia of the inner ear or
the vestibular nerve (Fig. 19.1). Nevertheless, a definitive clinical classification is not possible in individual patients with labyrinthine, vestibular nerve, or
vestibular nucleus lesions responsible for vertigo due
to AICA infarctions. A combination of peripheral and
central vestibular dysfunction is possible in patients

nocculus

paranocculus
Fig.19.1. Three zones of AICA supply. Zone 1 is supplied by the recurrent penetrating arteries (RPA) of AICA, zone 2 by the internal
auditory artery, and zone 3 by the terminal cerebellar branches of AICA. a Rostral pons at the level of the facial and abducens nuclei
(verticaf dotted fine represents the the mid-sagittalline). Zone 1A and zone 1B represent the arterial supply to areas supplied bya premeatal and postmeatal RPA. Often a single RPA, originating from the premeatal or postmeatal AICA, supplies all of zone 1. The crosshatched area represents the root entry zone of the facial and vestibulocochlear nerves. b Zone 2 represents the arterial supply to the
inner ear (adapted from Schuknecht 1974). c Cerebellum, anterior view. Zone 3 represents the arterial supply from the terminal cerebellar branches of AICA; ASe. anterior semicircular canal; AVA, anterior vestibular artery; CCA, common cochlear artery; HSC, horizontal
semicircular canal; IAA, interna I auditory artery; MCP, middle cerebellar peduncle; PSe. posterior semicircular canal; V, spinal trigeminal
tract and nucleus; VI, abducens nucleus; VII, facial nerve; VIII, vestibulocochlear nerve. (Adapted from Oas and Baloh 1992, with permission.)

Stroke and vertigo


with both vertigo and hearing loss (which supports
labyrinthine or eighth-nerve ischaemia).
Furthermore, it has been repeatedly demonstrated
that small demyelinating plaques (Brandt et al. 1986;
Dieterich and Bchele 1989) or lacunar infarctions
(Hopf 1987) at the root entry zone and/or the
vestibular nuclei can mimic vestibular neuritis
(p. 241). In such cases, rotational vertigo and spontaneous nystagmus are combined, and caloric
responses are abolished on the affected side. These
manifestations are either combined with masseter
paresis, as evidenced by the masseter reflex (Hopf
1987), or by ocular motor abnormalities such as
those of saccadic pursuit (Brandt et al. 1986).

Vertebral artery and posterior inferior


cerebellar artery
There are two major vascular syndromes of the
medulla oblongata: the medial and the lateral. The
medial medullary syndrome is characterised by the
triad of ipsilateral hypoglossal nerve palsy with contralateral hemiparesis and loss of deep sensation.
The lateral medullary infarction commonly presents
with Horner's syndrome, ataxia, alternating thermoanalgesia, nystagmus, vertigo, and hoarseness
(Gan and Noronha 1995). Further differentiations
try to separate small midlateral, dorsolateral, inferolateral, and inferodorsolateral infarcts (Vuilleumier
et al. 1995). A typical syndrome of occlusions of the
vertebral artery or the arteries arising from the vertebral artery (to course through the lateral
medullary fossa to supply the lateral medulla), and
in rare cases affecting the posterior inferior cerebellar artery (PI CA) or posterior spinal arteries, is
Wallenberg's syndrome (Wallenberg 1895; 1901), a
lateral medullary infarction that mostly includes the
medial, and sometimes the superior, vestibular
nucleus. A unilateral ischaemic lesion of the medial
(or superior) vestibular nucleus generally results in a
vestibular tone imbalance in the roll plane.
Signs and symptoms of vestibular dysfunction in
the roll plane (Chap. 10) can be detected by deviations from normal function. In the normal position
in the roll plane the graviceptive input from the
otoliths and semicircular canals aligns subjective
visual vertical (SVV) with gravitation al vertical and
adjusts axes of the eyes and the head horizontally. A
vestibular tone imbalance due to alesion should
result in a syndrome consisting of a perceptual tilt
(SVV tilt), head and body tilt, vertical misalignment
of the visual axes (skew deviation), and ocular torsion. The so-called ocular tilt reaction (OTR) com-

309

prises several of these features: the synkinesis of


head tilt, skew deviation, and ocular torsion
(Westheimer and Blair 1975) associated with a perceived tilt of the vertical (Brandt and Dieterich
1987).
There is convincing evidence that all these signs
and symptoms reflect vestibular dysfunction in the
roll plane. They may be found in combination or as
single components, with varying sensitivity at all
brainstem levels. A systematic study of 111 patients
with acute unilateral brainstem infarctions
(Dieterich and Brandt 1993a) revealed that pathological tilts of SVV (94%) and ocular torsion (83%)
are the most sensitive signs (Table 10.1). Skew deviation was found in one-third of these patients and a
complete OTR in one-fifth. The clinical evaluation of
vestibular function in roll therefore includes

detection of psychophysical adjustments of SVV,


determination of the vertical divergence of the
visual axes by use of prisms, and
determination of ocular torsion by means of fundus photographs (for methods see Dieterich and
Brandt 1993a).

Wallenberg's syndrome
All of the 36 patients with Wallenberg's syndrome we
tested exhibited significant tilts of the internal representation of the gravity vector, as indicated by deviation of SVV ipsiversive to the lesion. About one-third
of these patients had a complete OTR (Table 19.2,
Fig. 19.2; Dieterich and Brandt 1992); these were the
same patients exhibiting the most severe body
lateropulsion. Lateropulsion is a sensorimotor disturban ce that causes the body to deviate toward the
side of the lesion as if it is being pulled by some
external force (Bjerner and Silfverskiold 1968). OTR
in Wallenberg's syndrome is ipsiversive (i.e. ipsilateral ear and eye undermost). Quantitative measures
of postural sway by means of posturography demonstrate an increased diagonal sway from right forward
to left backward for right-side lesions and from left
forward to right backward for left-side lesions (Fig.
19.3). Body lateropulsion is correlated with SVV tilt
(i.e. the more pronounced the lateropulsion, the
greater the SVV tilt) (Fig. 19.4).
We hypothesise that deviation of SVV, lateropulsion of the body, skew deviation, and ocular torsion
of the eyes are the perceptual, postural, and ocular
motor consequences of a common lesion of the
vestibular pathways that subserve VOR in the roll
plane (Dieterich and Brandt 1993a,c). Most patients
have disconjugate ocular torsion, usually of the eye
ipsilateral to the brainstem lesion, which suggests

310

Vertigo

Table 19.2. Ocular motor abnormalities in 36 patients with Wallenberg's syndrome


Ocular motor abnormality

Frequency

(%)

Gaze-evoked nystagmus

35/36

(97)

14 horizontal bilateral (4c = i, Sc > i, Si > cl, 11 ipsiversive, 3 contraversive, 7 upwards

Lateropulsion ofthe closed eyes'

21/23

(91)

Evaluation of 23 EOGs: 18 ipsilateral (11 severe > 15,6 moderate 5- 15, 1 slight < 5),
3 contra lateral, 2 no lateropulsion

Spontaneous nystagmus

27/36

(75)

14 contraversive, 6 ipsiversive, 5 horizontal alternating, 2 upward, 4 no spontaneous


nystagmus (6 no registration)

Saccadic pursuit

26/36

(72)

17 horizontal bilateral, 5 ipsiversive, 3 contraversive, 7 up- + downwards, 4 upwards, 4


no deficit (6 no registration)

Skew deviation

16/36

(44)

Hypotropia of the ipsilateral eye

Ocular tilt reaction b

(33)
(25)

Triad of head tilt, skew deviation, and ocular torsion; all features ipsiversive

Central positional nystagmus

12/36
9/36
3/36

Horizontal gaze paresis

1/36

Dysmetria of saccades

(8.3)

Note

4 contralateral hypometria, 2 ipsilateral hypometria, 1 ipsi- + contra lateral, 2 slowness


Rotatory nystagmus in head-hanging position
Ipsiversive

' Lateropulsion = horizontal deviation of the eyes.


bOcular tilt reaction = triad of lateral head tilt, skew deviation, and ocular torsion in the direction of head tilt.
c> i =contralateral more than ipsilateral (c =contralateral; i = ipsilateral); EOG =electro-oculogram.
From Dieterich and Brandt (1992).

PERCEIVED TILT

thalamus, vest ibu lar cortex


~

EXCYCLOTROPIA

~SO
....,.

HYPERTROPlA

10

He

It
pe
?

otolithic input

"
8

@t

SR

IR

:-

""--

I
I

I
I

MLF

I
I
I

I
I

,
cerv ica l cord

HEAD TlLT

Fig.19.2. a Patient with a right lateral medullary infarction (Wallenberg's syndrome) presenting with an OTR to the right. OTR consists of ipsiversive head tilt of 20 (bottom), skew deviation of 4 (middle), and ocular torsion of the undermost eye of about 20 (excyclotropia; counterclockwise from the viewpoint of the observer; top), whereas the uppermost eye shows a normal position in roll (5
excyclotropia). b Hypothetical explanation of OTR due to lesions of the vertical semicircular canal pathways. Schematic drawing of the
three-neuron vestibulo-ocular reflex arc between the posterior semicircular canal and the extraocular eye muscles. An excitatory
ascending pathway is linked from the posterior semicircular canal to the ipsilateral superior oblique and the contra lateral inferior rectus muscles; an inhibitory ascending pathway is linked to the ipsilateral inferior oblique and the contra lateral superior rectus muscles
(Graf et al. 1983; Graf and Ezure 1986). Alesion of these pathways causes excyclotropia of the ipsilateral and hypertropia of the contralateral eye. AC, PC, HC, anterior, posterior, and horizontal semicircular canals; MLF, medial longitudinal fascicle; OS and 01, superior
oblique and inferior oblique muscles; RS and RI, superior and inferior rectus muscles; 111, oculomotor nucleus; IV, trochlear nucleus; VIII,
vestibular nucleus. (From Dieterich and Brandt 1992.)

Strake and vertigo

311
EYES CLOSED

EYES OPEN
A

normals

25'

:E
c:

10mm
I----<

.............

~
.:;;

20'

GI
"0

oEGI

lateropulsion 11

>

15'

:>'"
GI
.~

.S!.
.a

..........

..........

10'

.1.

t..1:1

5'

I i'i
I
O'

n .. 13

"_24

11

n.16

111

n.3

IV

Lateropulsion deviation
SWAY HISTOGRAM

Fig.19.3. Typical body sway histograms in two patients with


Wallenberg's syndrome during upright stance with the eyes open
and closed. Grade IIlateropulsion indicates considerable imbalance without falls; grade 111 lateropulsion represents falls with
eyes closed. Body sway increases considerably in patients with
Wallenberg's syndrome, and the normal dominance of the anteriorposterior (fore-aft) sway component shifts to an abnormal
dominance of the lateral sway component, which is diagonal in
grade 11 lateropulsion and becomes more lateral in grade 111
lateropulsion (From Dieterich and Brandt 1992.)

involvement of posterior semicircular canal pathways (Fig, 19.2). If anterior and posterior semicircular canal pathways are affected, OTR in Wallenberg's
syndrome manifests as binocular ocular torsion
(Fig. 10.7). In rare cases of monocular incyclotropia
of the uppermost eye, it can be assumed that only the
anterior semicircular canal pathways are involved.
Where is the most probable site of the lesion of
graviceptive pathways in roll in Wallenberg's syndrome? Projection of ischaemic lesions demonstrated by CT and MRI onto the appropriate sections of a
stereotaxic brainstem atlas (Olszewski and Baxter
1982) allows identification of the medial (and/or
superior?) vestibular nucleus as the critical vestibular structure (Fig. 19.5). The medial and/or superior
vestibular nuclei were included in the ischaemic

Fig.19.4. Tilts of subjective visual vertical (SW, tilt in degrees)


in 36 patients with infarctions of the lateral medulla oblongata.
Adjustments of SVV (first measurements in the acute stage) are
depicted in relation to the severity of body lateropulsion, as
classified in grades I to IV (abscissa). The dots represent single
measurements (in so me patients there were repeated measurements during the course of the disease); means and standard
deviations are also depicted. The more pronounced the lateropulsion, the greater the deviations of SVV (H test; p < 0.001).
SVV and lateropulsion are both ipsiversive to the side of the
lesion. (From Brandt and Dieterich 1993b.)

areas in cases of Wallenberg's syndrome with OTR.


The blood supply to these structures originates from
the PICA, from the branches of the vertebral artery,
or from the branches of the AICA.
Our studies on OTR, lateropulsion, and SVV
focused on static effects of vestibular dysfunction in
roll, which persist for days to weeks, during which
time they gradually and spontaneously subside.
Additional dynamic signs and symptoms consisting
of lateral rotational vertigo and torsional nystagmus
occur in the acute stage of infarction (Morrow and
Sharpe 1990; Lopez et al. 1992). Fast phases of rotational nystagmus are contraversive, whereas the slow
phases correspond in direction to the static deviation. The combination of static and dynamic effects
is not surprising if one considers the functional

312

Vertigo

XXIV
D27W30

E2M76

XVI
F34M30

F36M70

VIII m

.. vertebral

posterior spinal arteries er


PI CA

artery

Fig. 19.5. Unilateral pontomedullary infarctions causing


vestibular dysfunction in the roll plane, presenting either as OTR
or its components: head tilt, skew deviation, ocular torsion, and
tilt of SVV. Schematic representation of two transverse sections
(XVI, XXIV) from the stereotaxic brainstem atlas of Olszewski and
Baxter (1982), with typical lesioned areas in four patients with
unilateral medullary and pontine infarctions, as taken from MR
images and projected onto the appropriate transverse section.ln
the lateral medullary infarctions (Wallenberg's syndrome) - within the territories of the vertebral arteries, the PICA or the posterior spinal arteries - vestibular dysfunction with ipsiversive tilts
results from involvement of the medial vestibular nucleus (VII1m).
In pontine infarctions vestibular dysfunction in roll results either
from involvement of the superior vestibular nucleus (VIIls) or the
MLF within the territories of the AICA or the paramedian arteries
from the basilar artery, respectively (top). Occlusions of the AICA
may involve the superior vestibular nucleus and then cause
ipsiversive OTR, whereas MLF lesions cause contraversive OTR.

corroboration of vertical semicircular canals and


otolithic input based on neuronal convergence at the
second-order vestibular nuclei neurons (Angelaki et
al. 1993).
The cerebellum has, to a great degree, not yet
been considered in this chapter. Ischaemic lesions of

the vestibular vermis structures (uvula, nodulus)


and the nucleus fastigii can probably cause vertigo
or related vestibular syndromes, but it is unlikely
that infarctions of only the cerebellar hemispheres
do. The cerebellum is also involved in 3-dimensional
control of pursuit eye movements, which are organised in a way similar to that of the VOR. The signals
used for vertical smooth pursuit are, at some stage,
encoded in a semicircular canal-VOR coordinate
framework. The evidence for this was seen in a
pathological ocular torsion during vertical pursuit in
patients with cavernous angioma within the middle
cerebellar peduncle (FitzGibbon et al. 1996).
Moreover, experimentallesion of the nodulus in cats
has been reported to cause positional downbeat nystagmus (Fernandez et al. 1960; Fernandez and
Fredrickson 1964). Attempts to confirm this in
patients by Kattah et al. (1984) and Sakata et al.
(1987) were less convincing. Furthermore, our experience has shown that severe central positional vertigo
is usually induced by lesions located dorsolaterally
to the fourth ventricle (Chap. 18). As a general rule, it
is difficult to differentiate between brainstem and
cerebellar lesions in most clinical cases of infarctions, because the major infratentorial arte ries
supply both brainstem and cerebellum (Kase et al.
1993; Norrving 1995; Vuilleumier et al. 1995).

Basilar artery and paramedian


pontine and mesencephalic arteries
Acute circumscribed pontine lesions may occur with
infarcts and haemorrhages (Caplan and Han 1995)
or as pontine ischaemic rarefaction in elderly hypertensive patients with subcortical arteriosclerotic
encephalopathy-like pathology (Pullicino et al.
1995). The arterial system supplying the midbrain is
terminal with overlaps between arte rial territories
and individual variations (Hommel and Besson
1995).

Vestibular syndromes in roll plane


Unilateral ischaemic lesions of the pontomesencephalic vestibular pathways, which run along the
MLF, predominantly cause vestibular tone imbalance
in the roll plane. This is demonstrated by the frequency of SVV tilts, skew torsion, and OTR (often
combined with unilateral internuclear ophthalmoplegia) in affected patients. The direction of the
perceptual, ocular motor, and head tilt may be helpful for determining the level as weIl as the side of the

313

Stroke and vertigo

brainstem lesion. All tilts are ipsiversive (i.e. ipsilateral eye undermost) in cases of caudal pontomedullary lesions and contraversive (contralateral
eye undermost) in cases of rostral pontomesencephalic lesions (See also p. 181). These directionspecific findings of vestibular dysfunction in roll can
be explained by unilaterallesions of a "graviceptive"
pathway that crosses the midline at the upper
pontine level (Figs. 19.6 and 19.7; Brandt and
Dieterich 1993a), running along the MLF and reaching the interstitial nucleus of Cajal (INC), a wellknown integration cent re for eye-head co ordination
in roll (Anders on 1981; Fukushima 1987).
If the level of the brainstem damage is known
from the clinical syndrome, then an SVV tilt, OTR,
or skew torsion will indicate the side more severely
affected. If, however, the side of the damage is clear
from the clinical syndrome, then the level of the
brainstem damage will be indicated by the tilt direction of these signs (i.e. caudal with ipsiversive and
rostral with contraversive tilt). Thus, the diagnostic
topographie value of a static vestibular dysfunction
in roll is similar to that of cranial nerve lesions, but
the former is more sensitive. The directions of these
diagnostic rules will be reversed if the vestibular
dysfunction in roll is dynamic (torsional nystagmus). Torsional nystagmus ipsiversive to the side of
an MLF lesion has been described in combination
with unilateral internuclear ophthalmoplegia
(Dehaene et al. 1996). The MLF lesion could be
responsible for inactivating the ipsilateral INC,
which would result in contraversive ocular deviation
(slow phase). The presence of a corrective ipsiversive
quick phase implicates an intact rostral interstitial
nucleus of the MLF (riMLF) (Riordan-Eva et al.
1996). If the riMLF is lesioned by a circumscribed
infarction, then a seemingly paradoxical torsional
nystagmus may occur, a contralesionally beating torsional nystagmus (HeImchen et al. 1996). Thus,
lesions in both mesencephalic regions (INC or
riMLF) cause

(static) contraversive tonic ocular deviations in


roll, but
(dynamic) torsional nystagmus, whose direction
is either ipsiversive (INC) or contraversive
(riMLF).

One of the most striking, previously unreported,


findings was that all of our patients with skew deviations also had ocular torsion (OT) toward the undermost eye (p. 186). The direction-specific coincidence
of skew and OT is helpful for clinically differentiating between supranuclear (mesencephalic) skewand
nuclear or fascicular oculomotor or trochlear
palsies. A unilateral oculomotor or trochlear palsy

a
L

Fig. 19.6 Typicallesioned areas in four patients with unilateral


pontomedullary infarctions, showing ocular skew torsion sign to
the left (Ieft eye undermost) . lschaemic areas were taken from
MRI and projected onto the appropriate transverse sections of
the stereotaxic brainstem atlas of Olszewski and Baxter (1982).ln
medullary lesions involving the medial and superior vestibular
nuclei (Vilim in sections XIV and XVI, Vllis in section XXIV), the
skew torsion sign was ipsiversive (a-c). The ocular skew torsion
sign was contraversive if the medial longitudinal fascicle (Flom in
section XXIV) was involved, with the vestibular nuclei spared (d;
same section as in cl. Opposite directions of ocular skew torsion
signs in lesions on the same side at the pontine level (c,d) indicate crossing of graviceptive pathways mediating eye position in
the roll plane; in c, ipsilateral pathways were affected; in d, contralateral pathways were affected (From Brandt and Dieterich
1993a.)

Vertigo

314

causes OT of only the paretic eye, whereas mesencephalic skew can be identified most often by a
binocular, conjugated OT. Even in patients with
bilateral third or fourth nerve palsies, a skew can be
differentiated by the direction of OT, which is
dysconjugate in the peripheral nerve palsies (i.e.
bilaterally excyclotropic in bilateral trochlear palsies
and bilaterally incyclotropic in bilateral oculomotor
palsies). Therefore, we recommend determining
skew and OT (by fundus photographs) during
routine neuro-ophthalmological examination of a
patient with suspected brainstem dysfunction
(Dieterich and Brandt 1993b). Furthermore, dysfunction in the roll plane due to central brainstem
lesions can be reliably differentiated from OT and
SVV tilt caused by extraocular eye muscle paresis,
because the latter is not associated with SVV tilts
under binocular viewing conditions.

Vestibular syndromes in pitch plane


c

T pyr .
T PVr

Fig.19.7. Ischaemic areas in four patients with right pontine


and mesencephalic infarctions causing contraversive ocular skew
torsion sign to the left. Lesions indicated involvement of either
the medial longitudinal fascicle (Flom in sections XXVIII, XXX,
XXXVI) at different pontomesencephalic levels (a-c) or the rostral
midbrain tegmentum, including the interstitial nucleus of Cajal
(iC in section XXXVIII, d). The latter seems to be the most rostral
brainstem structure to elicit the ocular skew torsion sign (From
Brandt and Dieterich 1993a.)

A vestibular tone imbalance in the pitch plane


typically results from bilateral paramedian lesions
(p. 201; Chap. 11) of the vestibular pathways (This
may be helpful for early clinical diagnosis of, e.g.
basilar artery thrombosis). A tone imbalance in the
pitch plane can cause a downbeat nystagmus (p. 199)
if the lesion is located at the level of the vestibular
nuclei, affecting commissural fibres at the floor of
the fourth ventricle (Baloh and Spooner 1981).
Downbeat nystagmus in the primary gaze position,
especially in lateral gaze, is often accompanied by
oscillopsia and postural instability with increased
fore-aft body sway.
Upbeat nystagmus (p. 205) - the directional opposite of downbeat nystagmus - can be induced by two
separate and distinct brainstem lesions: either at the
ponto-medullary junction near the nucleus perihypoglossus or at the pontomesencephalic junction
(Fisher et al. 1983). Transitions between upbeat and
downbeat nystagmus are possible, depending on the
level of brainstem ischaemia. Pontomesencephalic
brainstem lesions, whether unilateral or bilateral,
may therefore cause a vestibular tone imbalance in
roll or pitch, but not in yaw. A tone imbalance in the
yaw plane (Chap. 12, p.215) indicates pontomedullary lesions near the root entry zone of the
eighth nerve or the vestibular nuclei.

Thalamic infarctions
The thalamus receives most of its blood supply from
four arterial pedicles that arise from the basilar

315

Stroke and vertigo

artery bifurcation, the posterior communicating


artery, and the proximal portion of the posterior
cerebral arteries (Barth et al. 1995; Fig. 19.10).
Unilateral thalamic infarctions may cause contralateral falling, astasia (Masdeu and Gorelick 1988), or
contralateral OIR (Halmagyi et al. 1990). We
measured SVV, OI, skew deviation, and lateral head
tilt in 35 patients with acute thalamic infarctions (14
paramedian, 17 posterolateral, 4 polar) (Iable 19.3,
Dieterich and Brandt 1993c) and in 5 patients with
mesodiencephalic haemorrhages, in order to determine the differential tonic effects on vestibular function in roll plane. Eight of 14 patients with
paramedian infarctions had complete OIR (Fig.
10.9) with contraversive head tilt, skew deviation,
OI, and SVV tilt. Projection of the infarcted areas
revealed on CI scans or MR images onto appropriate
transverse sections of cytoarchitectonic atlases
clearly showed that OIR is due to concurrent
ischaemia of the rostral midbrain tegmentum,
including the INC (Figs. 19.8 and 19.9).
Ihe blood supplies to the two regions may have a
common origin in the paramedian thalamic and
paramedian mesencephalic arteries (Fig. 19.10).
Ihus, OIR is not thalamic, and INC (and the rostral
interstitial nucleus of the MLF) is obviously the most
rostral brainstem structure that mediates eye-head
co ordination in roll (p. 181). Different clinical
patterns of these infarctions (involving third nerve
palsy and vertical gaze palsy) have been described
as paramedian thalamopeduncular infarction
(Iatemichi et al. 1992). Eleven of 17 patients with
posterolateral infarctions (Fig. 19.11) exhibited moderate but significant SVV tilts, which could be either
ipsiversive or contraversive. Ihalamic astasia with
preferred lateral body sway (Fig. 19.12) is the postural consequence of the perceptual tilt. In these cases
vestibular thalamic nuclei (Vim, Vce, Dc) were
involved, whereas infarctions were more ventro-

medial in the remaining cases. Polar infarctions did


not affect vestibular function in roll. Even if thalamic
infarction includes the vestibular subnuclei, it does
not result in spontaneous nystagmus or skew deviation.

Cortical infarctions
Animal studies have identified several distinct and
separate areas of the parietal and temporal cortex
which receive vestibular afferents, such as area 2v at
the tip of the intraparietal su1cus (Fredrickson et al.
1966; Schwarz and Fredrickson 1971; Bttner and
Buettner 1978), area 3aV (neck, trunk, and vestibular
region of area 3a) in the central su1cus (dkvist et al.
1974), the parietoinsular vestibular cortex (PIVC) at
the posterior end of the insula (Grsser et al. 1982;
1990a, b), and area 7 in the inferior parietallobule
(Faugier-Grimaud and Ventre 1989) (Chap. l3, Fig.
13.1, p. 220). Our knowledge about vestibular cortex
function in humans is less precise and is derived
mainly from stimulation experiments reported
anecdotally in the older literature. It is not always
possible to extrapolate from monkey species to the
human cortex, as Andersen and Gnadt (1989) have
demonstrated for Brodmann's area 7 in rhesus monkey and humans. Area 2v corresponds best to the
vestibular cortex as described by Foerster in 1936.
PIVC corresponds best to a region from which
Penfield and Jasper (1954) were able to induce
vestibular sensations by electrical stimulation with a
depth electrode within the Sylvian fissure, medial to
the primary acoustic cortex. Ihis region was found
to be activated during caloric vestibular stimulation:
there was a focal increase of cortical blood ftow
(Friberg et al. 1985).
What is the clinical significance of the vestibular

Table 19.3. Skew deviation, oeular torsion, subjeetive visual vertieal, and head tilt in different types of thalamie infaretions
Type of thalamie
infaretion

NO.of
patients

Skew
Na

Oeular torsion
Angle

SVV tilt

Na(B,M)

Head tilt

Na

Na

Angle

----~~~-

Paramedian
With OTR
Without OTR
Posterolateral
Polar

14
8
6
17
4

8
0
0
0

7.5

aNumber of patients with pathologieal findings


bAdditional third-nerve palsy.
(Ocular torsion was measured in 15 of 17 patients.
d SVV was measured in 16 of 17 patients.
Data are given as means in degrees.
From Dieterieh and Brandt (1993e).

8 (6B, 2Mb)
2b(2M)
Sc (3B, 2M)
0
B Binoeular.
M Monoeular.
Ilpsilateral eye.
CContra lateral eye.

9.0
9.0'b
2.8'
0

6.5
7.0'b
2.0'
0

8
1

lld
0

9.9
6.4'
4.3'
0

12.7"
7.3'
4.1'

8
1
0
0

10
5'
0

316

Vertigo

Fig.19.8. Typicallesioned areas in a paramedian thalamie infaretion with OTR a and a thalamie haemorrhage without OTR b, taken
from MR images and projeeted onto the appropriate transverse seetions of a stereotaxie thalamus atlas (Van Buren and Borke 1972:
9.7 mm and 0.9 mm above the anterior eommissure-posterior eommissure [AC-PC] line) (top and middle) and midbrain atlas
(Olszewski and Baxter 1982: plate XXXVIII) (bottom). The AC-PC interval is 25 mm. a The infaretion involves the rostral midbrain
tegmentum in the region of the interstitial nucleus of Cajal (iC = INC) and the adjaeent area of the rostral interstitial nucleus of the MLF
(riMLF). b The mesodieneephalie haemorrhage spares paramedian rostral midbrain struetures. (Abbreviations: Apr = nucleus anterior
prineipalis; Ce pe = nucleus eentralis parvoeellularis; Cma = anterior commissure; Cmp = posterior commissure; Cos = superior collieulus; Cun = nucleus euneiformis; Oe = nucleus dorsoeaudalis; 00 = nucleus dorso-oralis; Edy = nucleus endymalis; EW = EdingerWestphal nucieus; F = fornix; Fa = nucleus fascieularis; Gmpe = medial genieulate body, pars parvoeellularis; HI = nucleus habenularis
lateralis; Hm = nucleus habenularis medialis; iC = interstitial nucleus of Cajal [INC]; Icp = nucleus intraeapsularis; IIlpr = nucleus oculomotorius principalis; iLa = nucleus intralamellaris; Lem = mediallemniseus; Li = nucleus limitans; Lpo = nucleus lateropolaris; M =
nucleus medialis; NIII = oeulomotor nerve; Pf = nucleus parafaseieularis; PI = lateral pallidum; Pm = medial pallidum; Pt = nucleus
parataenialis; Pul = lateral pulvinar; Pum = medial pulvinar; Puo =oral pulvinar; Put = putamen; Pv = nucleus paraventrieularis hypothalami; R = retieular nuclei; Ru pe = red nucleus, pars parvoeellularis; SC = superior eollieulus; Smth = stria medullaris thalami; SN em =
substantia nigra, pars compaeta; TM = traet of Meynert; Tmth = mammillothalamie tract; Tte = central tegmental traet; Vee = nucleus
ventroeaudalis externus; Vci = nucleus ventroeaudalis internus; Vepc = nucleus ventrocaudalis parvoeellularis; Vim = nucleus ventrooralis intermedius; Voe = nucleus ventro-oralis externus; Voi = nucleus ventro-oralis internus (From Oieterieh and Brandt 1993c.)

Stroke and vertigo

317

a
Fig.19.9. Collective presentation of lesions (taken from MRls and projected onto the appropriate transverse thalamic and midbrain
sections) in seven patients with para median thalamic infarctions and two patients with mesodiencephalic haemorrhages. a In the
seven ischaemic patients who manifested with complete contraversive OTR, the rostral midbrain tegmentum was involved, including
the region of the INC (= iC) and the adjacent area of the riMLF. Black areas represent at least five overlaps. bin two patients without
clinical signs of OTR, the region of the INC appeared unaffected. (See Fig. 19.8 legend for abbreviations.) (From Dieterich and Brandt
1993c.)

cortex? Vertigo has long been recognised to be a


manifestation of epileptic seizures (Chap. 14)
(Foerster 1936; Penfield and Jasper 1954; Schneider
et al. 1968). This type of vertigo is secondary to unilateral focal discharges from the vestibular cortex,
whereas functional deficits of this area, such as those
caused by infarctions of the medial cerebral artery,
do not typically manifest with vertigo. Our study
required a reliable test for determining the functional
deficit of vestibular cortex lesions in patients with

acute ischaemic stroke. SVV was chosen as one


mode of testing vestibular function, because it provides a sensitive and direction-specific means for
measuring unilateral peripher al (Friedmann 1971;
Curthoys et al. 1991) and central (Dieterich and
Brandt 1993a) vestibular dysfunctions.
We systematically examined 71 patients presenting with infarctions of the middle, posterior and
anterior cerebral arteries for their visual ability to
adjust a test line to vertical. Skew deviation and OT

Vertigo

318
Internal
/ ' Carotid
Artery
_

/
Thalam icSubthalamic
Artenes ~

Thalamogeniculate
Artenes

Post.
/ . / Choroidal
./
Artenes

Basilar
Artery -

Post.
Comm,
Artery

Post
Cerebral
Artery

Fig. 19.10. A schematic representation of the blood supply to


the thalamus, including the polar artery, the thalamic-subthalamic
arteries, the thalamogeniculate arteries, and the posterior
choroidal arteries (modified after Barth et al. 1995). The thalamic
subthalamic (thalamoperforating) arteries often arise jointly
from the basilar (or posterior cerebral) artery and occlusion of the
proximal common vessel (arrow) causes a continuation of paramedian thalamic and rostral midbrain infarctions (with contra versive OTR, p. 319).

were also evaluated in order to complete perceptual


and ocular motor testing of vestibular function in
the roll plane. Infarcted areas shown on CI and MRI
were projected onto corresponding transverse
sections of the atlas of Duvernoy (1991) to identify
the critical structures involved. Ihe aim of the study
was to address the following three questions:

Are there (one or several) distinct supratentorial


areas, alesion of which causes a direction-specific
tilt of the SVV?
Is it possible to relate these areas to known
vestibular structures as identified in animal
experiments?
Are pathological tilts of SVV in supratentorial
lesions associated with skew deviation and
ocular torsion toward the perceptual tilt?

Our 71 patients with unilateral supratentorial infarctions were evaluated with respect to static vestibular
function in the roll plane; this also involved
determinations of the SVV, skew deviation and OI.
Infarctions in the territories of the posterior and
anterior cerebral arteries did not affect static
vestibular function in roll. Iwenty-three of 52
patients with infarctions in the middle cerebral
artery (MCA) territory showed significant, mostly
contraversive, pathological SVV tilts (Fig. 19.13,

Iable 19.4). Ihe overlapping area ofthese infarctions


centred on the posterior insula, which is homologous to the PIVC in the monkey. Although electrophysiological and cytoarchitectonic data in animals
demonstrate several multisensory areas rather than
a single primary vestibular cortex, the PIVC seems to
represent the integration centre of the multisensory
vestibular cortex areas within the parietal lobe (p. 220).

Cortical rotational vertigo


Unlike lesions of the vestibular nuclei or their brainstern pathways, an acute unilateral infarct of the
vestibular cortex does not usually manifest with vertigo. In vestibular epilepsy (Chap. 14), it is not the
functionalloss but rather the paroxysmal focal discharge that causes cortical vertigo. A unique case has
been described of a woman with embolie infarction
that occurred within the right middle cerebral artery
territory; she had non-epileptic rotational vertigo,
nausea, and unsteady gait that gradually resolved
within a week (Brandt et al. 1995). A well-demarcated
lesion of the posterior insula (homologous to the
parietoinsular vestibular cortex in monkeys) was
probably the cause. Ihis was supported by a contraversive tilt of perceived vertical and by the involvement of the adjacent auditory (Heschl's) cortex
lable 19.4. Pathological subjective visual vertical (SVV) tilt and ocular
torsion (OT) in supratentorial unilateral infarctions of the anterior cerebral
artery (ACA, n = 4), middle cerebral artery (MCA, n = 52), and posterior
cerebral artery (PCA, n = 15)
Vascular territory

SVVtilt
Static'

OT
Monocular'

ACA
MCA
Frontal branches
Temporal branches
Central branches
Parietal branches
Deep perforators

4
52
1
18
3
4
9

0
13 c, 1 i (6.2)
0
1c
5 c, 1 i (4)
2 c, 1 i
(3.5)
3 c (4)
2 c (3.4)
0
2 c (3.4)
0
0

0
3 c, 1 i (2.6)
0
0
1 c (6)

APIC
4
LSA
5
ACHA
10
PPIC
7
Temporal
3
Posterior border zone 7
PCA
15

0
1 c (2)
1i
0
0
0
0

APIC = anterior part of the interna I capsule; LSA = lenticulostriate arteries;


ACHA = anterior choroidal artery; PPIC = posterior part of the interna I
capsule; temporal = temporal region of the ACHA; n = number of patients;
c = contra lateral; i = ipsilateral.
, Figures in parentheses show the mean deviation in degrees obtained by
adding absolute values for static binocular SVV, but adding algebraicaily
for 01

319

Stroke and vertigo

Fig.19.11. Overlap areas of nine posterolateral thalamie infaretions that eaused either ipsiversive (i; n = 6) or eontraversive (e; n = 3)
tilts of SW. As ean be seen by eomparison with the transverse seetion of the middle thalamie level (9.7 mm above the AC-PC line), as
shown in panel a and the lower thalamie level (0.9 mm above the AC-PC lineL as shown in panel b, the overlap area involves the
thalamie nuclei Vee, De, Vim, Vci, and, less frequently, Voe, independently of the direetion of indueed tilt of the interna I representation
of gravity. Blaek areas represent six overlaps; hatehed areas represent five overlaps on the left side and two to three overlaps on the
right side. (See Fig. 19.8 legend for abbreviations.) (From Dieterieh and Brandt 1993c.)
normal
A

IOmm

..........

Posterolateral Thalamic Inlarction Left

Fig.19.12. Body sway histogram in a patient with left posterolateral thalamie infaretion du ring upright stanee with the eyes
open and elosed (bottom) . Note the inereased, predominantly
lateral sway ("thalamie astasia") whieh is assoeiated with a tilt of
pereeived vertieal.

M.J .d60

EYES OPEN

EYES CLOSED

320

Vertigo

+16

Fig.19.13. Collective presentation of infarcted areas taken fram MR images and projected onto the apprapriate transverse sections
of the atlas of Duvernoy (1991) for seven patients with clearly demarcated infarctions of the MCA wh ich caused significant contraversive tilts of perceived vertical. Overlapping areas of infarctions (7 of 7, in black) in the section + 16 mm above AC-PC line are centred at
the posterior part of the insula, involving the long insular gyrus with the adjacent short insular gyrus, the transverse temporal gyrus,
and the superior temporal gyrus. This area could represent the human homologue of the PIVC in monkey (see Fig. 13.1).

(Figs. 19.14 and 19.15). Dipole source analysis oflate


auditory evoked potentials (Fig. 19.15) revealed a
decreased amplitude of a dipole source in the right
insular cortex, whereas two other dipoles situated in
the temporal lobe showed symmetrie aetivity. The
latter demonstrate the close proximity of the
vestibular cortex to the auditory cortex.
The rarity of vertigo due to infarctions of the
middle cerebral arte ries is surprising. Our case is
distinctive, because vestibular and auditory cortex

areas were involved, whereas most of the neighbouring temporoparietal cortex was spared (Fig. 19.14).
One can speculate that the homologue of the parietoinsular vestibular cortex (Chap. 13) caused pereeptual tilt and rotational vertigo. However, one
cannot exclude the role of the homologue of area 7 at
the parietal cortex or a functional deficit of corticocortical connections between several vestibular
areas (Brandt et al. 1995).

Stroke and vertigo

321

c
Fig.19.14. MRI (after Gd-DTPA application) showing regional embolic infarction (increased signal intensity, see arrows) of the tem poral branches of the right middle cerebral artery. (a,b) Frontal sections at a level at about the posterior end of the insula (T2 -weighted
sequence on the left;T1-weighted sequence on the right) show the vertical extent ofthe lesion (arrows) that includes the vestibular
and the auditory cortex, respectively. c Transverse sections (Tl-weighted sequence) at a level about 16 mm above the AC-PC line
showing an insular lesion (arrows) involving the long insular gyrus and retroinsular regions as weil as the transverse temporal gyrus. d
Transverse section at a level above C (T1-weighted sequence) showing the horizontal extent of the posterior insular lesion and a second regional infarction of the right superficial parietal cortex (arrow) corresponding to the posterior upper part of multisensory area 7.
A cavum septum pellucidum is seen as a common variation. (From Brandt et al. 1995.)

322

Vertigo

~~~~r5PV
I

-100 0

400 ms

103 ms

R
143 ms

11~PV

R
I

400ms

Fig. 19.1 S. Late auditory evoked potential dipole source analysis. a Upper panel shows evoked potential to a 2000-Hz tone of 100ms duration. Electrode labels according to the International 10-20 system. Note a negative peak at approximately 100 ms after stimulus onset with different shape at left (T3, (3) and right (T4, (4) electrodes. b Schematic heads showing location of intracranial dipole
sources Nos.3 and 4 (left and right insular cortex) and corresponding time-varying activity. Note maximum activity of source No. 3 at
103 ms and decreased amplitude of contra lateral source. c At 143 ms, two more lateral sources in the temporal lobes are active with
symmetrical amplitudes. (From Brandt et al. 1995.)

References
Amarenco P (1991) The spectrum of cerebellar infarctions.
Neurology 41:973-979
Andersen RA, Gnadt JW (1989) Posterior parietal cortex. In:
Wurtz RH, Goldberg ME (eds) Reviews in oculomotor research.
vol. 3. The neurobiology of saccadic eye movements. Elsevier,
Amsterdam, pp 315-335
Anderson JH (1981) Ocular torsion in the cat after lesions of the
interstitial nucleus of Cajal. Ann NY Acad Sei 374:865-871
Angelaki DE, Bush GA, Perachio AA (1993) Two-dimensional spatio-temporal co ding of linear acceleration in vestibular nuclei
neurons. J Neurosei 13:1403-1417
Atkinson WJ (1949) The anterior inferior cerebellar artery. J
Neurol Neurosurg Psychiatry 12:137-151
Baloh RW (1992) Stroke and vertigo. Cerebrovasc Dis 2:3-10
Baloh RW (1996) Vestibular dis orders due to cerebrovascular disease. In: Baloh RW, Halmagyi CM (eds) Disorders of the
vestibular system. Oxford University Press, Oxford, pp 418-429
Baloh RW, Spooner JW (1981) Downbeat nystagmus: a type of
central vestibular nystagmus. Neurology 31 :304-310
Barth A, Bogousslavsky J, Caplan LR (1995) Thalamic infarcts and
haemorrhages. In: Bogousslavsky J, Caplan LR (eds) Stroke syndromes. Cambridge University Press, Cambridge, pp 276-283

Bjerner K, Silfverskiold BP (1968) Lateropulsion and imbalance in


Wallenberg's syndrome. Acta Neurol Scand 44:91-100
Brandt Th (1991) Man in motion. Historical and clinical aspects of
vestibular function. A review. Brain 114:2159-2174
Brandt Th, Dieterich M (1987) Pathological eye-head coordination in roll: tonic ocular tilt reaction in mesencephalic and
medullary lesions. Brain 110:649-666
Brandt Th, Dieterich M (1993a) Skew deviation with ocular torsion: a vestibular brainstem sign of topographie diagnostic
value. Ann Neurol 33:528-534
Brandt Th, Dieterich M (1993b) Vestibular falls. J Vestib Res 3:3-14
Brandt Th, Dieterich M (1995) Central vestibular disorders in roll,
pitch, and yaw planes: topographie diagnosis of brainstem disorders. Neuro-ophthalmologyI5:291-303
Brandt Th, Dieterich M, Bchele W (1986) Postural abnormalities
in central vestibular brain stern lesions. In: Bles W, Brandt Th
(eds) Disorders of posture and gait. Elsevier, Amsterdam, pp
141-156
Brandt Th, Dieterich M, Danek A (1994) Vestibular cortex lesions
affect the perception of verticality. Ann NeuroI35:403-412
Brandt Th, Btzel K, Yousry T, Dieterich M, Schulze S (1995)
Rotational vertigo in embolie stroke of the vestibular and auditory cortices. Neurology 45:42-44
Bttner U, Buettner UW (1978) Parietal cortex area 2 V neuronal
activity in the alert monkey during natural vestibular and optokinetic stimulation. Brain Res 153:392-397

Stroke and vertigo


Caplan LR, Han W (1995) Pontine haemorrhages and infarcts. In:
Bogousslavsky J Caplan LR (eds) Stroke syndromes. Cambridge
University Press, Cambridge, pp 324-335
Curthoys IS, Dai MJ, Halmagyi GM (1991) Human ocular torsional
position before and after unilateral vestibular neurectomy. Exp
Brain Res 85:218-225
Dehaene I, Casselman JW, D'Hooghe M, Van Zandijcke M (1996)
Unilateral internuclear ophthalmoplegia and ipsiversive torsional nystagmus. J NeuroI243:461-464
Dieterich M, Brandt Th (1992) Wallenberg's syndrome:
lateropulsion, cyclorotation and subjective visual vertical in 36
patients.Ann NeuroI31:399-408
Dieterich M, Brandt Th (l993a) Ocular torsion and tilt of subjective visual vertical are sensitive brainstem signs. Ann Neurol
33:392-399
Dieterich M, Brandt Th (I 993b ) Ocular torsion and perceived vertical in oculomotor, trochlear and abducens palsies. Brain
116:1095-1104
Dieterich M, Brandt Th (l993c) Thalamic infarctions: Differential
effects on vestibular function in the roll plane (35 patients).
Neurology 43:1732-1740
Dieterich M, Brandt Th (1995) Vestibular syndromes and vertigo.
In: Bogousslavsky J, Caplan LR (eds) Stroke syndromes.
Cambridge University Press, Cambridge, pp 80-90
Dieterich M, Bchele W (1989) MRI findings in lesions at the
entry zone of the eighth nerve. Acta Otolaryngol (Stockh) Suppl
468:385-389
Duvernoy HM (1991) The human brain. Surface, threedimensional sectional anatomy and MRI. Springer, Berlin
Heidelberg New York
Faugier-Grimaud S, Ventre J (1989) Anatomic connections of
inferior parietal cortex (area 7) with subcortical structures
related to vestibulo-ocular function in a monkey (Macaca fascicularis). J Comp NeuroI280:1-14
Fernandez C, Alzate R, Lindsay JR (1960) Experimental observation on postural nystagmus. 11. Lesions of the nodulus. Ann
Otol Rhinol LaryngoI69:94-114
Fernandez C, Fredrickson JM (1964) Experimental cerebellar
lesions and their effect on vestibular function. Acta Otolaryngol
(Stockh) 192:52-62
Fisher A, Gresty M, Chambers B, Rudge P (1983) Primary position
upbeating nystagmus: a variety of central positional nystagmus. Brain 106:949-964
FitzGibbon EJ, Calvert PC, Dieterich M, Brandt Th, Zee DS (1996)
Torsional nystagmus during vertical pursuit. J Neuro-ophthalmol
16:79-90
Foerster 0 (1936) Sensible kortikale Felder. In: Bumke 0, Foerster
o (eds) Handbuch der Neurologie, vol. VI. Springer, Berlin, pp
358-449
Fredrickson JM, Figge U, Scheid P, Kornhuber HH (1966)
Vestibular nerve projection to the cerebral cortex of the rhesus
monkey. Exp Brain Res 2:318-327
Friberg L, Olsen TS, Roland PE, Paulson OB, Lassen NA (1985)
Focal increase of blood flow in the cerebral cortex of man during vestibular stimulation. Brain 108:609-623
Friedmann G (1971) The influence of unilaterallabyrinthectomy
on orientation in space. Acta Otolaryngol (Stockh) 71:289-298
Fukushima K (1987) The interstitial nucleus of Cajal and its role
in the control of movements of head and eyes. Prog Neurobiol
29:107-192
Gan R, Noronha A (1995) The medullary vascular syndromes
revisited. J NeuroI242:195-202
Graf W, McCrea RA, Baker R (1983) Morphology of posterior
canal-related secondary vestibular neurons in rabbit and cat.
Exp Brain Res 52:125-138
GrafW, Ezure K (1986) Morphology ofvertical canal related second order vestibular neurons in the cat. Exp Brain Res 63:35-48
Grsser O-J, Pause M, Schreiter U (1982) Neuronal responses in
the parieto-insular vestibular cortex of alert Java monkeys

323
(Macaca fascicularis). In: Roucoux A, Crommelinck M (eds)
Physiological and pathological aspects of eye movements. Dr.
W. Junk, The Hague, Boston, London, pp 251-270
Grsser OJ, Pause M, Schreiter U (1990a) Localisation and
responses of neurons in the parieto-insular vestibular cortex of
awake monkeys (Macaca fascicularis). J PhysioI430:537-557
Grsser OJ, Pause M, Schreiter U (l990b) Vestibular neurons in
the parieto-insular cortex of monkeys (Macaca fascicularis):
visual and neck receptor responses. J PhysioI430:559-583
Halmagyi GM, Brandt Th, Dieterich M, Curthoys IS, Stark RJ, Hoyt
WF (1990) Tonic contraversive ocular tilt reaction due to unilateral meso-diencephalic lesion. Neurology 40:1503-1509
Heimchen C, Glasauer S, Bart! K, Bttner U (1996)
Contralesionally beating torsional nystagmus in a unilateral
rostral midbrain lesion. Neurology 47:482-486
Hinojosa R, Kohut RI (1990) Clinical diagnosis of anterior inferior
cerebellar thrombosis: autopsy and temporal bone histopathology
study.Ann Otol Rhinol LaryngoI90:261-271
Hommel M, Besson G (1995) Midbrain infarcts. In: Bogousslavsky
J, Caplan LR (eds) Stroke syndromes. Cambridge University
Press, Cambridge, pp 336-343
Hopf HC (1987) Vertigo and masseter paresis. A new local
brainstem syndrome probably of vascular origin. J Neurol
235:42-45
Kase CS, Norrving B, Levine SR, Babikian VL, Chodosh EH, Wolf
PA, Welch KMA (1993) Cerebellar Infarction: clinical and
anatomic observations in 66 cases. Stroke 24:76-83
Kattah JC, Kolsky MP, Luessenhof AJ (1984) Positional vertigo and
cerebellar vermis. Neurology 34:527-529
Kim HN, Kim YH, Park IY, Kim GR, Chung IH (1990) Variability of
the surgical anatomy of the neurovascular complex of the cerebellopontine angle. Ann Otol Rhinol LaryngoI90:288-296
Lopez L, Bronstein AM, Gresty MA, Rudge P, Du Boulay EPGH
(1992) Torsional nystagmus: a neuro-otological and MRI study
of 35 cases. Brain 115:1107-1124
Leigh RJ, Brandt Th (1993) Areevaluation of the vestibulo-ocular
reflex: new ideas of its purpose, properties, neural substrate,
and dis orders. Neurology 43: 1288-1295.
Masdeu JC, Gorelick PB (1988) Thalamic astasia: in ability to stand
after unilateral thalamic lesions. Ann NeuroI23:596-603
Morrow MJ, Sharpe JA (1990) Torsional nystagmus in the lateral
medullary syndrome. Ann NeuroI24:390-398
Norrving B (1995) Medullary infarcts and haemorrhages. In:
Bogousslavsky J, Caplan LR (eds) Stroke syndromes.
Cambridge University Press, Cambridge, pp 318-323
Oas JG, Baloh RW (1992) Vertigo and the anterior inferior cerebellar artery syndrome. Neurology 42:2274-2279
dkvist LM, Schwarz DWF, Fredrickson JM, Hassler R (1974)
Projection of the vestibular nerve to the area 3a arm field in the
squirrel monkey (Saimiri sciureus). Exp Brain Res 21:97-105
Olszewski I, Baxter D (1982) Cytoarchitecture of the human
brainstem. Karger, Basel
Penfield W, Jasper H (1954) Epilepsy and the functional anatomy
of the human brain. Litt!e Brown, Boston
Pullicino P, Ostrow P, Miller L, Snyder W, Munschauer F (1995)
Pontine ischemic rare faction. Ann NeuroI37:460-466
Riordan-Eva P, Faldon M, Bttner-Ennever JA, Gass A, Bronstein
AM, Gresty MA (1996) Abnormalities of torsional fast phase
eye movements in unilateral rostral midbrain disease.
Neurology 47:201-207
Sakata F, Ohtsu K, Shimura H, Sakai S (1987) Positional nystagmus
of benign paroxysmal type (BPPV) due to cerebellar vermis
lesions. Pseudo BPPV.Auris Nasus Larynx (Tokyo) 14:17-21
Schneider RC, Calhoun HD, Crosby EC (1968) Vertigo and rotational movement in cortical and subcorticallesions. J Neurol
Sei 6:493-516
Schuknecht HF (1974) Pathology of the ear. Harvard University
Press, Cambridge, Mass.

324
Schwarz DWF, Fredrickson JM (1971) Rhesus monkey vestibular
cortex: abimodal primary projection field. Science 172:280-281
Tatemichi TK, Steinke W, Duncan C, Bello JA, Odel JG, Behrens
MM, Hilal SK, Mohr JP (1992) Paramedian thalamopeduncular
infarction: Clinical syndromes and magnetic resonance imaging.Ann NeuroI32:162-171
Van Buren JM, Borke RC (1972) Variations and connections of the
human thalamus, vol. 2. Variations of the human diencephalon.
Springer, Berlin Heidelberg New York
Vuilleumier P, Bogousslavsky J, Regli F (1995) lnfarction of the

Vertigo
lower brainstem: Clinical, aetiological and MRI-topographical
correlations. Brain 118:1013-1025
Wallenberg A (1895) Acute Bulbraffection (Embolie der Art.
cerebelli. post. inf. sinistr.). Arch Psychiat Nervenkr 27:504-540
Wallenberg A (1901) Anatomischer Befund in einem als "Acute
Bulbraffection" (Embolie der Art. cerebelli. post. inf. sinistr.)
beschriebenen Falle. Arch Psychiat Nervenkr 34:923-959
Westheimer G, Blair SM (1975) The ocular tilt re action - a brainstem oculomotor routine. lnvest OphthalmoI14:833-839

Migraine and vertigo

Migraine is an episodic disorder that has headache


as its most prominent feature and occurs in various
combinations with autonomie, gastrointestinal and
neurologieal symptoms. Current estimates indieate
that 23 million Americans suffer from migraine
headaches with more than 11 million experiencing
significant headache-related dis ability (Stewart et al.
1992; Silberstein and Lipton 1994).The prevalence in
children is 2-4% (Bille 1962), with boys and girls
being equally affected; the prevalence after puberty
is ab out 14% for women and 6% for men (Pryse
Phillips et al. 1992; Stewart et al. 1992).
The incidence of vertigo in association with
migraine has been reported to range between 50%
and 70%, if one includes und er the heading of vertigo sensations of dizziness, light-headedness and
unsteadiness (Selby and Lance 1960; Kayan 1984;
Harker 1996). This surprisingly high incidence rate
does not, however, reflect the clinical importance of
vertigo in relation to other more characteristic and
distressing symptoms of migraine, since only onethird to one-half of these patients describe "true
vertigo." According to Kayan and Hood (1984) and
Mri and Meienberg (1993), vertigo appears to be
either the major symptom or a severe complication
of migraine in only 5-8% of cases.
To diagnose and treat this minority of migraine
sufferers effectively, it is important to know that vertigo attacks (with nystagmus and postural imbalance) may occur without accompanying headache
(Table 20.1) in children (benign paroxysmal vertigo
of childhood), adolescents and adults (basilar
migraine, benign recurrent vertigo).1t is essential to
obtain a thorough personal and family history in
order to explore any association of vertigo attacks
with headache or other typical features of migraine.
It is often difficult or impossible to differentiate
between benign paroxysmal vertigo of childhood,
basilar migraine and benign recurrent vertigo in
adults. All three conditions are special forms of
migraine, the diagnostic descriptions of which overlap to some degree. These forms may occur sequen325

Table 20.1.

Migraine and vertigo

Vertigo ottocks in migraine without typicol headache ("vestibular

migraine")
(hildren
Benign paroxysmal vertigo of childhood
(Benign paroxysmal torticollis in infancy)
Adults
Benign recurrent vertigo
Basilar migraine

Vertigo os 0 major symptom in migraine


Basilar migraine

Oizziness and motion sickness as facultative features in migraine


Migrainous aura
During headache attack
(In headache-free periods?)

Association of migraine with other vertigo disorders?


$usceptibility to motion sickness
Familial periodic ataxia
(Meniere's disease?)
(Benign paroxysmal positioning vertigo?)

tially or may alternate in the same individual, thus


making a strict, clinical differentiation irrelevant.
The term "vestibular migraine" (Dieterich and Brandt
1999) might be more appropriate in cases where
episodie vertigo is not associated with headache.
Apart from the distinct, usually short vertiginous
attacks, an increased sensitivity to motion - motion
sickness - is also a typical feature of migraine
(Cutrer and Baloh 1992). As part of a general sensory
hyperexcitability (photophobia, phonophobia,
osmophobia), this causes the patient to seek dark,
quiet rooms (Drummond 1986; Stewart et al. 1994).
Furthermore, motion sickness susceptibility (in the
headache-free interval) is significantly increased in
children with clearly established migraine (Bille
1962; Barabas et al. 1983) and in adults (Kuritzky et
al. 1981b; Kayan and Hood 1984).

326

Vertigo

Migraine

Table 20.2.

Migraine without aura (common migraine)

Diagnastic criteria
A. At least five attacks fulfilling criteria B-D

The formal criteria published by the International


Headache Society in 1988 on the classification and
diagnosis of headache dis orders, cranial neuralgia
and facial pain distinguish two major varieties:

B. Headache lasting 4 to 72 hours (untreated or unsuccessfully treated)

1. migraine without aura (formally known as com-

C. Headache has at least two of the following characteristics:


1. Unilaterallocation
2. Pulsating quality
3. Moderate or severe intensity (inhibits or prohibits daily activities)
4. Aggravated by walking stairs or similar routine physical activity

mon migraine),
2. migraine with aura (formally known as classic
migraine).

D. During headache, at least one of the following occurs:


1. Nausea and/or vomiting
2. Photophobia and phonophobia

Migraine with aura is subclassified into

migraine with typieal aura (homonymous visual


disturbance, unilateral numbness or weakness,
aphasia),
migraine with prolonged aura (lasting longer
than 60 min),
familial hemiplegic migraine,
basilar migraine,
migraine aura without headache,
migraine with acute-onset aura.

E. At least one of the following is present:


1. History and physical and neurological examinations do not
suggest an organic disorder.
2. History and/or physical and/or neurological examinations do
suggest such disorder, but it is ruled out by appropriate
investigations.
3. Such disorder is present, but migraine attacks do not occur for the
first time in dose temporal relation to the disorder.
From Headache Classification Committee of the International Headache
Society (1988) and Silberstein and Lipton (1994).

Table 20.3.

Migraine with aura (dassic migraine)

Other varieties of migraine include

Diagnastic criteria
A. At least two attacks fulfilling criterion B.

B. At least three of the following four characteristics are present:


1. One or more fully reversible aura symptoms occur, indicating brain
dysfunction.
2. At least one aura symptom develops gradually over more than 4
minutes, or two or more symptoms occur in succession.
3. No single aura symptom losts more thon 60 minutes.
4. Headache folio ws aura with a free interval of less than 60 minutes
(it also may begin before or simultaneously with the aura).

ophthalmoplegie migraine,
retinal migraine,
childhood periodie syndromes.

The clinical syndrome


The basic features of migraine are clearly refiected in
the diagnostic criteria presented in abbreviated form
in Tables 20.2 and 20.3. A typical migraine attack can
have five phases: the prodrome, the aura, the
headache itself, the headache termination and the
postdrome (Blau 1980).

Aetiology and pathomechanism


Genetic studies of twins, spouses and familial aggregations suggest a multifactorial mode of inheritance
in both migraine without aura and migraine with
aura (Russell and Oie sen 1993; Russel et al. 1996).
Genetic transmission of migraine has also been discussed as occurring either through a recessive gene
with high penetrance or through an autosomal dominant gene with reduced penetrance (Russel and
Olesen 1993; Raskin 1993). Familial hemiplegic
migraine is an autosomal dominant dis order that
has been mapped to chromosome 19 (Joutel et al.
1993). An allelism of familial hemiplegie migraine

C. History, physical examination, and, where appropriate, diagnostic tests


exdude a secondary cause.
From Headache Classification Committee of the International Headache
Society (1988) and Silberstein and Lipton (1994).

and cerebral autosomal dominant arteriopathy


(CADSIL) has been largely excluded (Dichgans et al.
1996). The age of onset of migraine attacks is under
30 years in 85% of the patients. Most of the affected
family members with familial hemiplegic migraine
have abnormal eye movements that manifest as
vestibulocerebellar dysfunction (Elliot et al. 1996).
The earlier view that migraine is a cerebral vasomotor disorder with cerebral arterial vasoconstriction causing the aura and extracranial vasodilatation
producing the pain, is no longer accepted. As attractive as this simple explanation appears to be, it is not
supported by the complex pathomechanism of
migraine currently under discussion. During the
aura phase, an inhibition of cortical neuronal activity

327

Migraine and vertigo

(spreading depression, Lauritzen 1987) obviously


causes the neurologieal symptoms. The thus induced
reduction in regional cerebral blood flow is indicated
by focal hypoperfusion in the posterior brain regions
(detected by the xenon-inhalation technique, Olesen
1991). The pattern of reduced blood flow does not
coincide with particular territories of cerebral blood
supply, and transcranial Doppler studies did not
show significant changes in blood flow velocities in
the middle cerebral and basilar arteries during aura
or headache phases of a migraine attack (Diener et
al. 1991; 1993). The heterogeneous results ofprevious studies of cerebral blood flow were due to the
use of different techniques and methodological
problems, for example, migraine attacks with and
without aura were not differentiated. Overall, however, no significant regional blood flow changes have
been reported during the headache phase of
migraine without aura (single-photon emission
computed tomography, Ferrari et al. 1995). Thus,
arterial vasoconstriction alone cannot be the underlying mechanism.
Instead, changes of neuronal activity in the dorsal
raphe nuc1eus and the locus coeruleus of the brainstern - centres of autonomie control of cerebral and
dural blood flow and antinociception (Goadsby and
Lance 1988) - were found to excite efferent neurons
in the trigeminal nuc1ei and thereby cause vasodilatation of dural arteries, plasma extravasation,
release of vasoactive substances (serotonin, calcitonin gene-related peptide, substance P), activation
of prostaglandins and degranulation of mast cells
(Moskowitz 1990). This cascade of effects results in
the perivascular aseptic inflammation of arte ries in
the dura mater, which are provided for by the afferent (C fibres) and efferent sensory fibres of the
trigeminal nerve (Edvinsson et al. 1988), thus inducing the typical headache via the central trigeminal
nucleus. The fact that the trigeminal nucleus extends
down to the cervical cord (C2) and its projections, to
neurons that receive afferent fibres from upper cervical roots may explain the localisation of pain in the
neck and occiput (Diener and Peatfield 1996), the
most frequent sites of pain in basilar migraine, but
not in migraine without aura. Coactivation of autonomic brainstem nuc1ei may result in accompanying
symptoms, such as polyurea, perspiration, arterial
hypotension, nausea and vomiting.
The important role played by the brainstem in the
pathophysiology is supported by findings of
positron-emission tomography (PET) (Weiller et al.
1995). PET in patients with unilateral migraine without aura exhibited a bilaterally increased blood flow
in cingulate, auditory and visual association cortices
during the attack and a persisting brainstem
activation during the symptom-free interval. This
significant, slightly lateralised activation in the
symptom-free interval was detected over several

planes of the brainstem from the pons to the midbrain and the periaqueductal grey matter and brainstern reticular formation. Maximal activation
seemed to occur in the dorsal raphe nuc1eus and the
locus coeruleus (Weiller et al. 1995).
All these findings point to a mechanism of neural
dysfunction triggered by certain brainstem centres.
The mechanism induces the release of vasoactive
substances from sensory neurons, which leads to
sterile inflammation of dural blood vessels with
extravasation of plasma proteins. This, in turn,
induces, on the one hand, headache and, on the
other, an increase in regional blood flow in the
brainstem, cingulate, auditory and visual association
cortices during migraine without aura (Weiller et al.
1995) or a decrease in regional blood flow in the
basilar artery territory and the temporal and occipital
cortices during basilar migraine (HMPAO-Spect,
Seto et al. 1994). Decreased regional blood flow in
migraine attacks, particularly in basilar migraine,
can occasionally cause ischaemic infarcts. The prevalence of migrainous infarcts was reported to be as
high as 4% by Sturzenegger and Meienberg (1985).
The annual incidence of cerebral infarction associated
with migraine has been estimated at 3.4 per 100 000
in a community-based stroke study (Heinrich et al.
1986). The infarcts mostly occur in the occipital
base; vascular risk factors are uncommon and prognosis is generally good (Hoekstra-van Dalen et al.
1996).

Management
Pharmacological treatment of migraine may be
acute (abortive, symptomatie) or preventive (prophylactic). Patients who experience frequent and
severe headache often require both approaches
(Silberstein and Lipton 1994; Diener and Peatfield
1996). One or more of the following medications are
used in the acute treatment (Table 20.4) of
headaches of different severities: analgesics,
antiemetics, ergots, or serotonin 1D-receptor agonists.
Preventive treatment (Table 20.5) inc1udes betareceptor blockers, calcium antagonists, serotonin
antagonists, antidepressants and anticonvulsants.
Prophylaxis should be considered whenever the
attacks cannot be treated satisfactorily (Diener and
Peatfield 1996):

more than two to three migraine attacks per month


migraine attacks lasting for more than 48 hours
migraine attacks that the patient describes as
unbearable
complicated forms of migraine with neurologieal
symptoms that persist for more than 7 days.

328

Vertigo

Table 20.4. Therapy of acute migraine attacks in the adult


Substance
Antiemetics'
Metoclopramide

Domperidone
Analgesics b
Acetylsalicylic
acid
Paracetamol
Ibuprofen
Others
Ergotamine
Dihydroergotamine
(DHE)
Sumatriptan

Dose

Indication

Side effects

Contraindications

10-20 mg po, 20 mg
rectal, 10 mg iv, im

Improves nausea and


vomiting, absorption
of analgesics

Children < 12 years

10-20 mg po

See above

Extrapyramidal
system,oculogyric
crisis (treatment
biperiden iv)

500 mg-1 9 po, as


effervescent tablet,
500 mg-lg iv
500 mg-1 9 po,
rectal suppository
100-400 mg po

Moderate headache

Gastric pain,
bronchospasm,
haemorrhages

GI ulcer, asthma, tinnitus

Moderate headache

See aspirin

See aspirin

2 mg po, rectal,
max/attack = 4 mg,
max/month = 16 mg
1 mg sc or iv, is not
sufficiently absorbed orally
100 mg pO,6 mg sc

Severe headache

Nausea, vomiting,
tightness ofthe
ehest, headaehe
See ergotamine

Peripheral artery disease,


coronary heart disease,
hypertension
See ergotamine

Heat sensation,
fatigue, dizziness,
ehest tightness

See ergotamine

Moderate headache

Severe headache
Severe headache, early
vomiting and diarrhea

Liver disease

'If not effeetive after 10 min apply analgesie.


bin mild to average attacks apply antiemetie and analgesie; in severe attaeks apply antiemetie and other substanee.
From Diener and Peatfield (1996).

Table20.5. Migraine prophylaxis


Substanee

Dose/day

Remarks and
meehanism of action

Side effeets

Contraindieations

----------

Beta-receptor blockers
Metoprolol

Propranolol
Calcium antagonists
Flunarizine
Serotonin antagonists
Pizotifen

Fatigue, hypotonia,
disturbances of sleep,
bronchospasm, bradyeardia

AV-bloek, bradyeardia,
asthma, diabetes

Initially 5 mg, later 10 mg Long half-life


in men, 5 mg in women

Fatigue, weight gain, depression,


Parkinsonism, tremor

Depression,obesity,
extrapyramidal disorder

3 x 0.5 mg

Fatigue, weight gain, antieholinergic Pregnancy, glaucoma,


CHD, PVD, hypertension,
effeets, headaehe, nausea, muscle
hepatic disease
pain, dizziness, edema

Initially 50 mg, later


150-200 mg

Beta-l-selective

Initially 40 mg, later


160-200 mg

Non-seleetive

Lisuride

3 x 0.025 mg

Dopamine agonist

Methysergide

2-8 mg

Not> 3 months
(retroperitoneal fibrosis)

Probably effeetive substanees


2-3
Dihydroergotamine

x 1.5 mg

Not predietably absorbed


after oral intake

Nausea, drug-indueed headaehe

See acute treatment

Naproxen

2-3

x 250 mg

Inhibitor of prostagiandin
synthesis

GI problems, neutropenia count

GI uleer, thromboeytopenia,
asthma, pregnaney,
hepatie disease

Aeetylsalieylie acid

50-300 mg

Inhibitor of prostagiandin
synthesis

GI problems, asthma

Valproie acid

600-1200 mg

Amitriptyline

25-150 mg

Asthma, tinnitus, weight


gain, alopecia, tremor
Antidepressant

Dry mouth, arterial hypotension

Note: Effeetive substanees are in the order of therapeutie ehoiee. CHD = coronary heart disease.
From Diener and Peatfield (1996).

329

Migraine and vertigo

Basilar migraine (BM) and "vestibular


migraine"
BM is a relevant differential diagnosis for episodic
vertigo and accounts for 6% of patients seen in our
dizziness unit (Table 2.1). Diagnosis of BM can be
difficult in monosymptomatic cases, elderly patients
and referrals who have had a single attack without
history of familial or individual migraine.
Six relevant clinical studies have been conducted
involving larger sampies of patients with BM (34-91
patients):
1. BM was originally described by Bickerstaff
(1961a) to be a typical condition of the adolescent girl.
2. The study by Kayan and Hood (1984) particularly
emphasised the presence of visual, ocular motor
and auditory abnormalities during the interval
between two attacks, and found onset of vertigo
at the age of 40 years and older in more than 50%
of their patients.
3. The study by Sturzenegger and Meienberg (1985)
demonstrated that BM occurs throughout life
(from 10 to 62 years) with varying symptomatology and a surprising high incidence of impaired
consciousness (77%).
4. The study by Olsson (1991) documented a fluctuating low frequency sensorineural hearing loss
in more than half of his patients and frequent
changes in hearing immediately before migraine
attacks.
5. The study by Cutrer and Baloh (1992) took a different approach, correlating migraine and dizziness independently of BM.
6. The study by Dieterich and Brandt (1999) delineated an unusual but not infrequent subgroup of
BM ("vestibular migraine"), which manifests
throughout life and is characterised by frequent
monosymptomatic audiovestibular or ocular
motor attacks without headache. The suspected
diagnosis was confirmed in these patients only
by effective medical treatment, e.g. prophylaxis
with beta blockers.
As valuable as these studies are, they provide quite
heterogeneous information on the typical age of
onset, key symptoms, proof of diagnosis and association with headache. The different studies reflect the
authors' individual interests, a fact which creates
some confusion about the clinical entity, the spectrum of BM and - most relevant to the present discussion - its relation to vertigo. This is further
complicated by describing benign paroxysmal vertigo

of childhood and benign recurrent vertigo in adults


as migraine equivalents without headache. Doubts
are justified about the usefulness of such a clinical
separation of these recurrent types of episodic vertigo from BM or "vestibular migraine" as the disease
entity.

The clinical syndrome


Premonitory migraine symptoms with dysfunction
of the brainstem or the occipital cortex can be traced
to branches of the vertebral and basilar artaries.
Occurring mainly in adolescent girls, they were first
described as basilar artery migraine by Bickerstaff in
1961. His original description of the clinical syndrome is still valid although incomplete (Bickerstaff
1961a):
Age. Of the 34 patients who had this syndrome to
greater or lesser degree, all but 2 were under the
age of 23. All were under 35.
Sex. 26 of the 34 patients were adolescent girls.
Family history. A clear-cut history of migraine in
close relatives was obtained in 28 cases.
Onset. The first symptom was usually visual; in
some this consisted of totalloss of vision, and in
others of positive visual manifestations throughout both visual fields, usually so intense as to
obscure vision. This was followed by vertigo, ataxia
of gait, dysarthria, and occasionally tinnitus, not
necessarily in that order, and the ataxia was not
necessarily associated with vertigo. Sensory manifestations consisted in tingling or numbness in
the periphery of both hands and both feet, and
sometimes around both lips and on both sides of
the tongue.
Course. These symptoms lasted from 2 min to a
maximum of 45 min and then subsided rapidly; if
there had been complete loss of vision, this disappeared more gradually (5 min) through aperiod
of greying of vision.
The headache. In each case the disappearance of
the premonitory symptoms was accompanied by
severe throbbing headache, usually in the occipital region, and often accompanied by vomiting.
This appeared to relieve the headache, which
finally ceased when the patient siept.
Frequency of attacks. These attacks were infrequent, but in the girls were strikingly related to
menstruation. In between these attacks many
patients had more classical migrainous episodes.
Over the years there is a tendency for this type of

Vertigo

330

attack to cease and to be replaced by the more


common variety.
Since there is no pathognomonic test or sign that
provides a definitive diagnosis, Bickerstaff (1961a)
emphasised that the following, taken in conjunction,
are diagnostic: the sequence of events, the occurrence of other more common migrainous attacks at
other times, the return to complete neurological normality in a short time, the headache following the
attack being more commonly occipital than hemicranial and a positive family history. Subsequendy,
impairment or loss of consciousness was described
as a facultative symptom of the syndrome
(Bickerstaff 1961b; Lees and Watkins 1963). In our
experience, transient global amnesia (TGA) mayaIso
occur, probably secondary to involvement of thalamic
and temporal lobe arteries originating from the posterior circulation. The prevalence of migraine in
patients with TGA is ab out 25% and, thus, significantly higher than in control groups (Crowell et al.
1984). Diffusion-weighted MRI of TGA patients
shows elevated signal intensity in the left or both
medial temporal lobes, which is compatible with
spreading depression of neural activity (Strupp et al.
1998).
According to the International Headache Society
(1988), the attacks must contain two or more of the
following features in order to establish a diagnosis of
BM:

visual symptoms in both temporal and nasal


fields of both eyes
dysarthria
vertigo
tinnitus
decreased hearing
diplopia
ataxia
bilateral paraesthesias
bilateral paresis
decreased level of consciousness

The occurrence of these signs and symptoms reflect


dysfunction of the occipitallobes, brainstem, cerebellum and/or cranial nerves. Their specificity, however, as indicators of BM is less reliable.
From our own retrospective study on 90 patients
finally diagnosed as having BM (or "vestibular
migraine"), we learned to use medical treatment
trials and central ocular motor signs in the symptomfree period as additional indicators to establish the
diagnosis (Dieterich and Brandt 1999). If vestibular
or ocular motor dysfunction was the key symptom,
diagnosis of BM was based, first, on the his tory of at
least three attacks and second, on one of the following four typical constellations (A-D):

(A)

recurrent attacks of nonvestibular neurological


deficits attributable to the vertebrobasilar territory
associated headache
individual his tory of migraine

(B)

recurrent attacks of vestibular and/or ocular


motor dysfunction
associated headache
individual history of migraine

(C)

recurrent attacks of vestibular and/or ocular


motor dysfunction
associated headache
efficacy of medical (migraine) treatment

(D)

recurrent vestibular and/or ocular motor dysfunction


without associated headache
efficacy of medical (migraine) treatment

This categorisation in A-D allowed us to define constellations based on the minimal features necessary
to assess the diagnosis. Our patients showed the features of at least one of these constellations. Many
presented with features of more than one category,
especially when observed over a longer time.
Efficacy of medical treatment of migraine me ans
that either acute attacks were suppressed by ergotamines and/or the frequency of the attacks was significantly reduced by preventive medication with
beta-blockers (metoprolol) or flunarizine.
To illustrate how the paucity of recurrent uniform
symptoms observed in BM is often misleading, perhaps one case will suffice (Dieterich and Brandt
1999).
Case report
A 50-year-old woman with arterial hypertension
presented after three attacks of double vision,
mydriasis, and ptosis due to right-sided oculomotor palsy combined with slight bilateral internuclear ophthalmoplegia (INO) and vertical and
horizontal saccadic pursuit. Two episodes lasted
for 20 minutes, the last episode - the first time we
saw the patient - for 3 days without associated
headache. Transient ocular motor deficits developed within 5-10 minutes; residual deficits were
found in the symptom-free interval after the third
attack in the form of bilateral dissociated gazeevoked nystagmus, and vertical and horizontal
saccadic pursuit. There was no individual or
familial history of migraine. Normal angiography

Migraine and vertigo

331

excluded vascular diseases such as aneurysm,


arterial disseetion, other malformations, and atherosclerotic plaques. Cardiac diagnostics, including transesophageal echocardiography, indicated
that cardiac embolism was unlikely.
We first misinterpreted the condition as transient ischaemic attacks, and began an antiplatelet
therapy first with acetylsalicylic acid (100 mg
daily) for aperiod of 6 months, and then with
ticlopidine for another 4 months. Despite this
treatment attacks recurred with unaltered severity
and a frequency of 3-6 per month, each lasting 30
minutes to several hours. Twelve months after the
initial examination a few attacks were accompanied by moderate upper cervical and bilateral
occipital pain. In view of the ongoing monosymptomatic attacks, the exclusion of vascular disorders, and the prompt relief of the attack by
subcutaneous or oral sumatriptan (a 5-HTagonist), BM was suspected. Preventive medication was tried with the beta-blocker metoprolol
succinate (Beloc-Zok). Neither monotherapy with
metoprolol (1.5 tablets per day) nor with pizotifen
alleviated these ophthalmoplegie attacks. Now the
attacks were accompanied by severe occipital
headache. Episodic headache occurred as separate
attacks 4-6 times per week without associated
ocular motor or neurological deficits. Oral
sumatriptan was also able to interrupt the
headaches. A combination of the beta-blocker
metoprolol (95 mg daily) and flunarizine (2 x 5
mg daily) was then used; it dramatically improved
both kinds of attacks, episodic ophthalmoplegia
and headache. With this medication the patient
became symptom-free after 2 months. The medication was continued for 9 months before flunarizine was gradually reduced. The patient, who is

still taking metoprolol medication for arterial


hypertension (I tablet per day), has remained
symptom-free now for 3 years.
Ophthalmoplegie migraine is a well-known form of
migraine with aura. Neuro-ophthalmological evaluation of our patient dearly showed that the functional
deficits were due to central mesencephalic dysfunction, including nuclear and supranuclear ocular
motor structures rather than the peripheral third
nerve. This makes ophthalmoplegie migraine a subtype of BM. The confusing aspect of this case was the
simultaneous absence of headache in the first 12
months of the monosymptomatic episodes but evidence of slight ocular motor deficits (such as INO
and saccadic pursuit) in the symptom-free interval.
Having observed this pattern now in several other
patients, we learned that it is compatible with BM.
Data of 90 patients with BM were analysed with
respect to the following relevant aspects (Dieterich
and Brandt 1999):
Age of onset. The first BM attack occurred
between ages 10-82 years in females and 7-72 years
in males; the female-male ratio was 1.5:1. Mean age
of females was 38 years with a plateau in the third to
fifth decades (Fig. 20.1). Mean age of males was 42
years with a peak in the fifth decade. The very first
BM attack can manifest in males and females as late
as in the sixties without recall of a prior experience
of migraine attacks.
Association of BM with headache. In 68% of most
attacks BM was associated with moderate, (rarely
severe) bilateral pain in the neck and occiput, a holocephalic tension-type headache, or less frequently a
unilateral, temporal, frontal headache. One-third of
the patients (32%) could not recall headache.
Association with other forms of migraine. BM was

n =90

26
24

c:
C1l

-...

20

1ij

16

12

C1l
..Cl

a.

~ cl'n = 36
D <fn=54

:::I

c:

c,

0
0

10

20

30

40

50

60

70

80

90

years of age
Fig.20.1. Age of first manifestation of episodic vertigo as key symptom of basilar migraine ("vestibular migraine"). (From Dieterich
and Brandt 1999.)

332

Vertigo

the only manifestation of migraine in 48%, whereas


52% reported other forms of migraine.
Vestibular symptomatology. The prevailing type of
vertigo in BM was rotational (Table 20.6), sometimes
combined with positional vertigo. The second most
common type was to-and-fro vertigo with postural
imbalance and unsteadiness. It was striking that 78%
of patients could be classified according to their
episodic vertigo as belonging to a monosymptomatic audiovestibular group (auditory symptoms,
16%).
Vestibular and ocular motor deficits in the symptom-free interval. A surprisingly high percentage of
patients (66%) exhibited central vestibular or nonvestibular ocular motor signs independent of the
attacks (Table 20.7). These deficits were moderate
and showed some variation during subsequent evaluations between the attacks, but they did not completely disappear even with effective prophylactic
migraine treatment. Only 26% of the patients (Table
20.7) had no abnormal ocular motor function in the
symptom-free interval.
Diagnosis of BM with episodic vertigo ("vestibular
migraine")

A synopsis of all studies of BM with vertigo and ocular motor disorders not only confirms the
heterogeneous spectrum of neurological deficits
attributable to the vertebrobasilar artery territory,
Table 20.6. Characteristics* of vertigo during basilar migraine attacks in
90 patients
Rotational vertigo

To-and-fro vertigo

Duration of single attacks


seconds
< 10 minutes
10-60 minutes
hours
1-4days

70
7
15
8
27
13

34
2
3
6
14
9

Frequency of attacks
X perday
X perweek
X permonth
X peryear

62
1
17
26
18

33
4
11

Positional vertigo

14

n=82

n=8
Positional vertigo only
Dizziness and gait ataxia only
Ocular motor disturbances without vertigo

11
7

1
1
6

*Patients may report more than one feature in single or subsequent


attacks
From Dieterich and Brandt (1999).

Table 20.7. Basilar migraine: central ocular motor signs in the


symptom-free interval

Normal
Congenital strabismus
Congenital nystagmus

23
5
3

25.6
5.5

Central ocular motor signs'

59
43
20
24
10
10
8
3
2
3
1

65.6
48
22
27

Saccadic pursuit, vertical


Saccadic pursuit, horizontal
Gaze-evoked nystagmus
Spontaneous nystagmus (>5 o /S)b
Positional nystagmus
Dissociated gaze-evoked nystagmus
Downbeat nystagmus
Upbeat nystagmus
Impaired fixation suppression ofVOR
Nuclear oculomotor palsy

3.3

11
11
9
3.3
2.2
3.3
1.1

'Multiple quotations possible


bin combination with other central ocular motor signs and/or a directional
preponderance in caloric testing
From Dieterich and Brandt (1999).

but further provides us with vestibular and ocular


motor features hitherto not considered typical for
BM (Table 20.8). Attacks with predominantly monosymptomatic vertigo without other severe neurological deficits can be attributed to the pontomedullary
region, whereas attacks with more complex deficits,
including nuclear and supranuclear ocular motor
signs, seem to involve the brainstem more diffusely
and affect consciousness more frequently. The following six findings on BM with episodic vertigo
("vestibular migraine") seem to be clinically most
relevant for the appropriate diagnosis and treatment
of patients:
1. BM with episodic vertigo manifests throughout

life with a me an age of about 40 years.


2. One-third of the patients (independently of age)
do not complain of headache associated with
aura deficits.
3. Monosymptomatic (audio), vestibular, or ocular
motor attacks recur.
4. The duration of vestibular and ocular motor dysfunction varies, ranging from seconds to hours
or even days, most frequently lasting a few minutes or several hours.
5. Ocular motor abnormalities pers ist during the
interval of the attack.
6. Pharmacological (prophylactic) treatment is
effective.
The unusuallarge time frame of the duration is in
accordance with the bimodal distribution for duration, described earlier, of minutes to 2 hours and

333

Migraine and vertigo

Table 20.8. Basilar migraine with episodic vertigo as the key symptom
("vestibular migraine")

C/inicalsyndrome
- Episodes of rotational or (less frequent) to-and-fro vertigo lasting from
seconds to days with a duration of a few minutes or, most frequently,
several hours
-Increased sensitivity to motion du ring the attack ("motion sickness")
and increased susceptibility to motion sickness in between the attacks
-In 33% of patients episodic vertigo is not associated with headache
-In 66% of patients episodic vertigo and ocular motor deficits occur
without associated (vertebrobasilar) neurological deficits
-In 33% vertigo is associated with visual symptoms, dysarthria, tinnitus,
decreased hearing, diplopia, ataxia, bilateral paraesthesia, bilateral
paresis, or decreased level of consicousness
Incidence/age/sex
- 5-8% of migrainous population, female:male ratio in adults = 1.5 : 1,
first manifestation throughout life with a mean age of about 40 years
Aetiology/pathomechanism
- Genetic transmission, either through recessive gene with high
penetrance or autosomal dominant gene with reduced penetrance
- The pathomechanism involves migrainous spreading depression or
hypoperfusion
Cause/prognosis
- As in other forms of migraine
- Different forms of migraine attacks occur in alternation or sequentially
in the same individual
Management
- Acute migraine attack
Analgesics (acetylsalicylic acid, paracetamol, ergotamine,
sumatriptan)
Antiemetics (metoclopramide, domperidone)
- Prophylaxis for migraine attacks
Beta-receptor blockers (metoprolol, propranolol)
Calcium antagonists (flunarizine)
Serotonin antagonists (pizotifin)
Differential diagnosis
- Vestibular paroxysmia (disabling positional vertigo), Meniere's disease,
transient ischaemic vertebrobasilar attacks, central vestibular ataxia,
familial episodic ataxia, vestibular epilepsy
Benign paroxysmal vertigo in childhood and recurrent vertigo in adults
are subtypes of basilar migraine

longer than 24 hours (Cutrer and Baloh 1992;


Lempert et al. 1993). Other studies reported that BM
deficits typically last only a few seconds to minutes
(Kayan and Hood 1984) or 30 to 60 minutes
(Bickerstaff 1961a,b). The current diagnostic criteria
as defined by the International Headache Society in
1988 unfortunately do not include short (seconds to
minutes) aura deficits.
The variability of the duration of attacks, however,
bears a considerable impact on the differential diagnosis, which includes very short events such as
vestibular paroxysmia (p. 117), paroxysmal brainstern attacks, or longer-Iasting attacks as occur in
Meniere's disease or transient ischaemic attacks.
Many of the ocular motor findings in BM during the

attack and particularly in the symptom-free interval


are incompatible with Meniere's disease. The combination of vestibular and auditory symptoms - as is
typical for Meniere's disease or vestibular paroxysmia - was in our study comparatively rare in BM
(16%) (Dieterich and Brandt 1999).
The course of BM shows how difficult it is to
establish the correct diagnosis only by careful history
taking and clinical evaluation. In addition recourse
to medical treatment trials is necessary to support
one or the other differential diagnosis. For example,
efficacy of ergotamines or sumatriptan during an
attack or prophylactic beta-blockers support the
diagnosis of BM, whereas efficacy of carbamazepine
supports the diagnosis of vestibular paroxysmia.
Efficacy of betahistine and diuretics may support
dia gnosis of Meniere's disease; however, this is less
reliable. Sometimes it is necessary to try several prophylactic agents or combinations consecutively
before recurrent BM attacks can be sufficiently suppressed and thus the diagnosis hardened (see Case
report, p. 330). Differential diagnosis with vertebrobasilar transient ischaemic attacks may be difficult or impossible for a single or a few attacks, since
headache may or may not occur in both conditions.
Isolated vertigo, however, is an uncommon manifestation of vertebrobasilar ischaemia (Estol et al.
1996), and recurrent purely monosymptomatic
audiovestibular or ocular motor attacks support the
diagnosis of BM (Dieterich and Brandt 1999). Preexisting cerebral vascular disorders, earlier strokes
and severe cardiovascular risk factors may support
the diagnosis of transient ischaemia, but even then a
prophylactic medication with antiplatelet substances
or anticoagulants may be necessary to confirm the
diagnosis. Sometimes ongoing monosymptomatic
attacks despite prophylactic antiplatelet drugs may
finally lead to an appropriate diagnosis of BM.

Pathomechanisms of vertigo, motion sickness


and ocular motor deficits
Origin of vertigo in migraine

Vertigo may arise from alesion or inadequate stimulation of peripheral or central vestibular structures
encompassing brainstem, thalamus, or the parietoinsular vestibular cortex. The regular association of
vertigo in BM with central ocular motility disturbances (Table 20.7) strongly suggests that both arise
from brainstem dysfunction involving the vestibular
nuclei in cases of rotational vertigo and spontaneous
nystagmus (Brandt and Dieterich 1995). Transient
brainstem dysfunction was shown by significant
alterations of brainstem auditory evoked potentials

Vertigo

334

in the pontomesencephalic brainstem (prolonged


interwave III - V interval) during BM attacks; this
condition returned to normal with clinical recovery
(Jamada et al. 1986; Ganji et al. 1993). Some other
authors found a surprisingly high frequency of
peripheral audiovestibular deficits, which were related
to the attacks in more than 29% (Kayan and Hood
1984) and also occurred in the symptom-free interval in 21% (Cutrer and Baloh 1992). They argued
that a unilateral hyporesponsiveness on caloric testing proved peripheral dysfunction in some patients
diagnosed as having BM (Kayan and Hood 1984;
Eviatar 1981). However, there are other examples of
lacunar infarctions and small demyelinating plaques
involving the root entry zone of the eighth nerve
and/or the vestibular nuclei whieh can mimie
peripheral vestibular disease by combining rotation al
vertigo and spontaneous nystagmus and diminishing calorie responses on the affected side (Hopf
1987; Dieterich and Bchele 1989). Furthermore, differences in the frequency of peripheral signs and
symptoms in BM can in part be explained by the different definitions of a "vestibular canal paresis" on
caloric testing: this was based on side asymmetries
of either more than 22% (Cutrer and Baloh 1992) or
more than 39%, as in our study. Nevertheless, temporary or permanent peripheral vestibular and auditory deficits can be found in some BM patients,
which by themselves are indistinguishable from
signs and symptoms of peripherallabyrinthine disease (Harker and Rassekh 1987). In these patients
the involvement of the labyrinth itself may be
causative secondarily to the BM-induced decrease in
regional blood fiow of the inner ear. The inner ear is
supplied by the anterior inferior cerebellar artery
(AI CA) through the internal auditory artery
(branching in common cochlear and anterior
vestibular arteries). The AICA arises from the basilar
artery and also supplies the anterolateral pons, the
middle cerebellar peduncle and the cerebellar fiocculus (Kim et al. 1990). Adefinitive clinical classification of the lesion site, however, is not possible in
individual patients with labyrinthine, vestibular
nerve, or vestibular nucleus dysfunction.
Moreover, the theoretical possibility of combining
central and peripheral audiovestibular dysfunction
in a BM attack appears to be plausible in some cases.
Such a combination is supported by the distribution
of neuro-otological signs during the attack; these
signs were classified to be of central origin in 14%,
peripheral origin in 7.5%, and indeterminate origin
in 17.5% (Kayan and Hood 1984). If the aforementioned infratentorial anatomical structures are
affected in BM, the disease can mimic other peripherallabyrinthine diseases such as Meniere's disease,
thus posing problems for the differential diagnosis.

This would explain the significant difference in the


prevalence of migraine in "classic" (22%) and
vestibular (81 %) Meniere's disease and underline the
different aetiologies for the two conditions (Rassekh
and Harker 1992). A correct diagnosis can only be
established by the time course, family history, additional central ocular motor signs and, especially, the
appropriate treatment.
It is most likely that vertigo in BM is due to a
functional vestibular tone imbalance caused by an
asymmetrie activation or deactivation of bilateral
vestibular neuronal activity during the attack. This
may predominantly involve vestibular brainstem
structures extending from the pontomedullary
vestibular nuclei (inducing rotational vertigo) to the
rostral midbrain tegmentum (inducing ocular motor
dysfunction). It corresponds best to the asymmetric
activation of the entire pontomesencephalic brainstern structures found in PET studies (WeiHer et al.
1995). A distressing vertigo (in particular, the rotational type) is atypieal for supratentorial vestibular
thalamic and cortical lesions unless it occurs in
unique ischaemic events (Brandt et al. 1995) or in
vestibular epilepsy arising from the superior temporallobe (Foerster 1936; Penfield and Jasper 1954).

Motion sickness-Iike symptoms


It is striking that patients with both BM and

migraine without aura complain of motion sicknesslike sensations (up to 50%), i.e. that normal body
motion provokes dizziness and nausea during the
attack. This led Cutrer and Baloh (1992) to speculate
that the pathomechanism of migraine may be
induced by hyperexcitability of vestibular receptors.
In fact, it has been shown in animal experiments that
neuropeptide release (e.g. calcitonin gene-related
peptide) - elicited by trigeminal afferents - can
increase spontaneous neuronal activity of the inner
ear vestibular receptors (Adams et al. 1987). If the
sensitivity of the vestibular system increases - in
particular, when the hypersensitivity is asymmetric
-, susceptibility to motion sickness should also
increase even under normally adapted motion stimulation (Cutrer and Baloh 1992). In our opinion, this
increasing sensitivity must not be peripheral, for it
would cause similar effects if it were central, e.g. at
brainstem level. Hypersensitivity to other sensory
stimuli such as auditory, olfactory and visual stimuli
are well-recognised symptoms of migraine and one
of the basic diagnostie features according to the
Headache Classification Committee of the
International Headache Society. Phono- and photophobia were also attributed to significantly
increased blood fiow in auditory and visual association cortices during the migraine attack (measured

335

Migraine and vertigo

by PET scans with the eyes closed, WeiHer et al.


1995). Thus, motion sickness-like symptoms can be
interpreted as the consequence of neuronal hyperactivity and the associated extensive, increased blood
ftow in the brainstem, especially in the vestibular
nuclei of the pontomedullary brainstem.
Ocular motor deficits in the symptom-free interval
indicate permanent brainstem or cerebellar
dysfunction

The surprisingly high incidence (65%) of pathological ocular motor findings of central origin in the
symptom-free interval in our patients agrees with
the report by Kayan and Hood (1984), who - in reference to the attack - described "findings indicative
of definite dysfunction of the vestibular and/or
cochlear systems in 77.5% of their patients, half with
central and half with peripheral pathology" (18.8%
central, 28.8% peripheral, 30% inconclusive). Their
study, however, provides no data on neuro-otological
disturbances in the symptom-free interval. The high
frequency (21 %) of persisting peripheral vestibular
deficits in the study by Cutrer and Baloh (1992) was
not confirmed in our study (8.3%). Objectively measurable electronystagmographic abnormalities
(57-80%) were stressed earlier by Drsteler (1975),
Toglia et al. (1981) and Eviatar (1981). Kayan and
Hood (1984) speculatively interpreted these findings: "most migrainous complications are caused by
ischaemia of sufficient severity to produce infarction
of nervous tissue occurring during the vasoconstrictive phase of the attack". This simple explanation is
no longer accepted (see p. 326).
Ocular motor abnormalities in the symptom-free
interval may reftect subtle continuous neuronal dysfunction in certain brainstem nuclei as it is shown
by the persisting brainstem activation on PET. These
ocular motor deficits are not typicallate (cumulative) ischaemic sequelae of BM, for they can be
observed in young patients presenting after their
first attacks. This was reported earlier in the literature (Eviatar 1981), and we were able to confirm this
observation in our younger patients. The combination of recurrent episodic vertigo and ocular motor
abnormalities in the symptom-free interval recalls
familial episodic ataxias (Chap. 25, p. 365), which
have been recently identified as channelopathies.
Episodic ataxia type 1 is due to amissense point
mutation in the potassium channel gene (KCNAl)
on chromosome 12p13 (Griggs and Nutt 1995).
Patients with another form of episodic ataxia (type
2) generally show interictal nystagmus and headache
(Gancher and Nutt 1986) and often develop progressive ataxia and dysarthria with cerebellar vermian
atrophy. This latter disorder was recently localised to

chromosome 19p (Vahedi et al. 1995) and was combined with familial hemiplegic migraine, which is
also linked to chromosome 19 (Vahedi et al., 1995).
In episodic ataxia type 2, an abnormally elevated pH
level in the cerebellum was measured with nuclear
magnetic resonance spectroscopy. This is corrected
- as in some other channelopathies - by acetazolamide (Bain et al. 1992). By analogy one could speculate that brainstem deficits in the symptom-free
interval of BM reftect an electrophysiological neuronal membrane instability in one of the ion channels (sodium, calcium, chloride, potassium, or
transmitter-gated channel), which is known to be the
pathomechanism in a number of other inherited disorders with episodic neuronal dysfunction (Ptacek
et al. 1994; Browne et al. 1994).

Benign paroxysmal vertigo in childhood


The syndrome of benign paroxysmal vertigo (Table
20.9), first described by Basser in 1964, has its on set
in young children, usually between 1 and 4 years of
age, and is not a rare disorder (its prevalence rate is
2.6%, Abu-Arafeh and Russell 1995). A cardinal
symptom is the presence of sudden brief vertigo
attacks with postural imbalance and gait ataxia. The
frightened child, unable to move, stand or sit without support, clutches a person or an object.
Associated signs and symptoms are pallor, sweating,

Table 20.9.

Benign paroxysmal vertigo of childhood

Clinical syndrome
Sudden transient attacks (duration seconds to minutes) of incapacitating
vertigo, postural imbalance and gait ataxia, associated with nystagmus,
pallor, nausea, vomiting but no headache or impairment of consciousness
Incidence/age/sex
One of the most frequent vertigo syndromes in childhood (prevalence
rate: 2.6%), age of onset usually between 1 and 4 years, rarely above 10
years, equal sex distribution, personal or family history of migraine in twothirds of cases
Pathomechanism
"Migraine equivalent"
Course/prognosis
Attacks of varying frequency (about 1O/year) sometimes in clusters, cease
spontaneously within months to years, may convert into other forms of
migraine
Management
Attacks brief and self-limiting (no treatment necessaryJ, no controlled
study available on effective prophylaxis
Differential diagnosis
(Basilar migraineJ, vestibular epilepsy, central or peripheral vestibular
paroxysmia, familial episodic ataxia, perilymph fistula, Meniere's disease,
psychogenic vertigo.

Vertigo

336

nausea, vomiting, nystagmus and divergent strabismus, but not impairment of consciousness. The latter is of particular importance for differentiation of
this condition from vestibular epilepsy. The attacks
usually last from seconds to minutes, rarely for
hours. They occur when standing, sitting or lying;
the frequency varies from several times per week to
once a year. There are no precipitating factors, and
the attacks are not associated with headache. The
disease is benign in character, and there is spontaneous remission within a few years.
Whereas Basser (1964) wrongly described the
syndrome as a variety of vestibular neuritis, Fenichel
(1967) reported on two siblings in whom the attacks
progressively converted into dassic migraine. The
dose relationship between benign paroxysmal vertigo and migraine is striking (Koenigsberger et al.
1970; Eviatar and Eviatar 1974; Watson and Steele
1974; Dunn and Snyder 1976; Koehler 1980; EegOlofsson et al. 1982), because of the benign paroxysmal character, the frequently positive family history
of migraine, and the combination or replacement of
benign paroxysmal vertigo by other forms of
migraine (Lanzi et al. 1994). Transitions are possible
from this apparently distinct entity to benign paroxysmal torticollis in infancy as weIl as to basilar
migraine. In fact, three of the eight young children
presented by Golden and French (1975) as having
basilar migraine in childhood exhibited a symptomatology which also fits the diagnostic criteria of
benign paroxysmal vertigo. The majority of their
young patients were girls, a familiar pattern in basilar migraine, but Basser (1964) and Eeg-Olofson et
al. (1982) emphasise that benignparoxysmal vertigo
occurs equally frequently in girls and boys.
To date there is little information on the origin
and site of the vertigo in benign paroxysmal vertigo
in childhood, and there is no convincing evidence
that it should be located in the posterior temporal
cortex as supposed by Eviatar (1981). He interpreted
the direction or preponderance in caloric testing as
indicative of a temporal lobe dysfunction. This view
cannot be maintained, since distressing rotational
and to-and-fro vertigo are extremely rare conditions
in temporal lobe dysfunction (Brandt et al. 1995). It
is quite likely that the brainstem is involved, thus
making this condition a variant of BM.
As helpful as it may be to use a particular label for
this condition in childhood, it should not be treated
as a distinct and separate dinical entity. Three children and several adults in our study on BM
(Dieterich and Brandt 1999) - seen after the first
attacks - could have been dassified as having benign
paroxysmal vertigo of childhood or benign recurrent vertigo in adults. With longer observation

periods, however, an increasing number of them


developed additional signs and symptoms of BM.
Consequently, these patients would have been given
two different subsequent diagnoses, although the
underlying mechanisms remained the same.
Differential diagnoses should indude vestibular
epilepsy (MRI, EEG, loss of consciousness), posterior
fossa tumours (MRI, progressive course) and psychogenic dis orders (older children).
Koenigsberger et al. (1970) reported that the
caloric responses in these children are permanently
reduced. Subsequent investigators (Koehler 1980;
Eeg-Olofsson et al. 1982; Mira et al. 1984), however,
have found normal caloric responses. Some patients
present with a positional nystagmus in the vertigofree interval (Eeg-Olofsson et al. 1982). Headache
provocation tests with nitroglycerin, histamine and
fenfiuramine were positive in 9 of 10 patients, and in
4 patients induced a typical vertiginous attack (Mira
et al. 1984).
Management. According to the presumed
mechanism of a "migraine equivalent", one could
argue that management of the attack and prophylactic treatment should be the same as for other forms
of migraine (p. 327), but no controlled study yet
available gives an estimate of the efficacy of such a
regimen. On the other hand, attacks are mostly so
brief (lasting from seconds to minutes) that subsequent administration of drugs is unnecessary.
Prophylaxis, therefore, may only be required in
selected cases, depending on the frequency and
severity of the attacks before they eventually cease
spontaneously within some months or years.
Uncontrolled treatment of single cases of the disorder showed positive responses to dimenhydrinate
(Koenigsberger et al. 1970) and antiserotonergic
drugs (Mira et al. 1984).

Benign paroxysmal torticollis in infaney


Snyder (1969) described a paroxysmal dis order of
torticollis with its onset in the first year of life and
spontaneous recovery before 5 years of age.
Recurrence, duration, and a benign course with dose
association with migraine resemble paroxysmal vertigo of childhood (Sanner and Bergstrm 1979), in
which torticollis may be a feature of the symptom
(Eeg-Olofsson et al. 1982). Whereas benign paroxysmal vertigo in childhood may refiect dysfunction of
the pontomedullary vestibular nu dei, paroxysmal
torticollis may refiect dysfunction of the mesencephalic tegmentum near the zona incerta and interstitial nudeus of Cajal.

Migraine and vertigo

Benign recurrent vertigo


In adult patients presenting with recurrent, brief
episodes of vertigo and normal caloric responses in
the absence of cochlear signs, basilar migraine (see
p. 329) or (more specific?) benign recurrent vertigo,
a subtype of BM, should be suspected. The latter
syndrome was first described by Slater (1979) in
seven patients, later confirmed by Moretti et al.
(1980) in five patients, and most probably accounts for
a proportion of the 86 patients reported by Leliever
and Barber (1981) to have recurrent vestibulopathy
of unknown cause.
According to Slater (1979) and Moretti et al.
(1980), the syndrome typically consists of sudden
intense vertigo (not necessarily rotatory) and postural imbalance lasting for minutes or hours, rarely
days, associated sometimes with nausea and spontaneous nystagmus or positional nystagmus. The
vertigo subsides gradually and becomes prevailingly
positional in character. Patients do not complain of
tinnitus, hearing loss, fullness of the ear, or
headache. Caloric testing is normal in the vertigofree interval. The age of onset ranges from 7 (thus,
meeting the diagnostic criteria of benign paroxysmal vertigo of childhood, p. 376) to 55 years; females
are predominantly affected, with vertigo episodes
frequently recurring at menstruation. Most patients
are migraine sufferers, in whom benign recurrent
vertigo occurs at different times than the migraine
attacks, or they have a family his tory of migraine.
Other features in common with migraine include
precipitation by alcohol, lack of sleep, emotional
stress and other migraine triggers. The frequency of
the attacks shows large inter- and intraindividual
variation, from once daily to twice a year.
When it occurs in children, benign recurrent vertigo is indistinguishable from benign paroxysmal
vertigo of childhood (p. 376), although the mean
duration of the attacks is longer. It is our opinion
that benign recurrent vertigo in adults is not a
separate entity but (as discussed on p. 329) a monosymptomatic form of basilar migraine with transitions to and associated with other forms of migraine
in the affected individuals. Furthermore, basilar
migraine, benign paroxysmal vertigo in childhood,
and benign recurrent vertigo respond to the same
treatment. Recognising, however, that monosymptomatic attacks of audiovestibular and/or ocular
motor dysfunctions of varying durations are typical
manifestations of migraine with aura is more relevant than a nosological controversy about subtypes
of BM. To ensure that patients presenting with
monosymptomatic migrainous episodic vertigo
receive effective treatment, one might propose the
use of the more appropriate term "vestibular
migraine" (Dieterich and Brandt 1999).

337

Dizziness and vertigo as facultative symptoms


in migraine apart from BM
Vertigo and disturbance of balance have been
reported to occur as part of the aura preceding
migrainous headache. However, they appear most
commonly during the headache phase (Kayan and
Hood 1984) apart from BM. The complaints during
the headache phase in migraine are mostly oflightheadedness, dizziness and unsteadiness, but they
also include rotational illusions. It is striking that
these symptoms are worsened by vestibular stimulation produced by head movements or postural
changes (see Motion sickness in migraine, p. 334).
The link between motion sickness and migraine was
based on the high incidence of a history of motion
sickness in migraine sufferers: between 25% and
60% (Barabas et al. 1983; Childs and Sweetnam 1961;
Kayan 1973, 1984; Pearce 1971; Kuritzky et al.
1981a,b). In contrast, the incidence in normal populations or unselected patients with tension
headaches was less than 25% (Kayan and Hood
1984).
Migraine sufferers mayaiso develop vertigo during headache-free periods. The incidence of vertigo
not associated with headache is considerably higher
than in a non-migrainous population (Kuritzky et al.
1981b; Kayan and Hood 1984). Early descriptions of
episodes of vertigo in migraine without headache
stern from Heveroch (1925) and Symonds (1926).
Vertigo attacks may occur with and without a following headache (in alternating sequence) (Watson
and Steele 1974).

Association of migraine with other vertigo


disorders?
As summarised by Kayan (1984), a link has been
repeatedly asserted to exist between various associated disorders and migraine:
1. Children: bilious attacks, paroxysmal abdomi-

nal pain, motion sickness, benign paroxysmal


vertigo, benign paroxysmal torticollis in infancy
2. Adults and/or children: cyclical oedema,
epilepsy, paroxysmal tachycardia, pseudo-angina,
bronchial asthma, arterial hypertension,
benign recurrent vertigo, benign paroxysmal
positioning vertigo, Meniere's disease.
A positive link between migraine and Meniere's
disease (p. 88) was first mentioned by Meniere himself (1861) and supported later by Atkinson (1943,
1944), Hinchcliffe (1967) and Kayan and Hood

338

(1984). Finally, benign paroxysmal positioning vertigo (p. 264) was reported to be frequently associated
with migraine (Schiller and Hedberg 1960; Kayan
and Hood 1984). However, the possibility of a common causal relationship of migraine attacks,
endolymphatic hydrops and canalolithiasis remains
obscure and unconvincing.
Those disorders among the foregoing which are
related to vertigo, such as benign paroxysmal vertigo
of childhood (p. 376), benign paroxysmal torticollis
of infancy, (p. 336) and benign recurrent vertigo
(p. 337) are believed to represent "migraine equivalents" - a "migraine without headache" as described
by Whitty (1967). Functional hypotheses on motion
sickness and Meniere's disease (p. 88) have been
presented that propose a common vasomotor mechanism. Even more speculative hypotheses have been
based on the common features in certain personality
traits of these patients. The latter cannot explain the
surprisingly high incidence of these conditions in
migraine patients, which seems too high to be
coincidental.

References
Abu-Arafeh I, Russell G (1995) Paroxysmal vertigo as a migraine
equivalent in children: a population-based study. Cephalalgia
15:22-25
Adams I, Mroz E, Sewell W (1987) A possible neurotransmitter
role for CRGF in a hair-cell sensory organ. Brain Res
419:347-351
Atkinson M (1943) Meniere's syndrome and migraine: observations on common relationship.Ann Intern Med 18:797-808
Atkinson M (1944) Meniere's syndrome and certain related conditions. Eye Ear Nose Throat Monthly 23:436-445
Bain PG, O'Brien MD, Keevil SF, Porter DA (1992) Familial periodic
cerebellar ataxia: a problem of cerebellar intracellular pH
homeostasis.Ann NeuroI31:147-154
Barabas G, Matthews WS, Ferrari M (1983) Childhood migraine
and motion sickness. Pediatrics 72:188-190
Basser LS (1964) Benign paroxysmal vertigo of childhood. Brain
87:141
BickerstaffER (1961a) Basilar artery migraine. Lancet i:15-18
Bickerstaff ER (1961b) Impairment of consciousness in migraine.
Lancet ii: 1057 -1059
Bille B (1962) Migraine in school children. Acta Pediatr Scand 51
(Suppl):I-136
Blau JN (1980) Migraine prodromes separated from the aura:
complete migraine. Br Med J 281:658-660
Brandt Th, Dieterich M (1995) Central vestibular syndromes in
roll, pitch and yaw planes. Topographic diagnosis of brainstem
dis orders. Neuro-ophthalmology 15:291-303
Brandt T, Btzel K, Yousry T, Dieterich M, Schulze S (1995)
Rotational vertigo in embolic stroke of the vestibular and auditory cortices. Neurology 45:42-44
Browne DL, Gancher ST, Nutt JG, Brunter P, Smith EA, Kramer P,
Litt M (1994) Episodic ataxial myokymia syndrome is associated
with point mutations in the human potassium channel gene,
KCNA. Nature Genet 8:136-140

Vertigo
Childs AJ, Sweetnam MT (1961) A study of 140 cases of migraine.
Br Industr Med 18:234-236
Crowell GF, Stump DA, Biller I, McHenry LC, Toole JF (1984) The
transient global amnesia-migraine connection. Arch Neurol
41:75-79
Cutrer FM, Baloh RW (1992) Migraine associated dizziness.
Headache 32:300-304
Dichgans M, Mayer M, Mller-Myhsok B, Straube A, Gasser T
(1996) Identification of a key recombinant narrows the
CADASIL gene region to 8cM and argues against allelism of
CADASIL and familial hemiplegic migraine. Genomics
32:151-154
Diener HC, Peters C, Rudizo M, Noe A, Dichgans I, Haux R,
Ehrmann R, Tfelt-Hansen P (1991) Action of ergotamine, fiunarizine and sumatriptan on cerebral blood fiow velo city in
normal subjects and patients with migraine. J Neurol
238:245-250
Diener HC, Limmroth V, May A, Laurich F, Auerbach P, Wosnitza
G, Eppe T (1993) Changes of cerebral blood fiow velocity after
treatment of migraine with sumatriptan or placebo.
Cephalalgia 13:14-17
Diener HC, Peatfield RC (1996) Migraine. In: Brandt Th, Caplan L,
Dichgans J, Diener HC, Kennard C (eds) Neurological disorders: course and treatment. Academic Press, San Diego, pp 1-15
Dieterich M, Bchele W (1989) MRI findings in lesions at the
entry zone of the eighth nerve. Acta Otolaryngol (Stockh)
Supp!. 468:385-389
Dieterich M, Brandt Th (1999) Episodic vertigo related to
migraine (90 cases): vestibular migraine? J Neurol (in press)
Drummond PD (1986) A quantitative assessment of photophobia
in migraine and tension headache. Headache 26:460-469
Dunn D, Snyder CH (1976) Benign paroxysmal vertigo of childhood.Am J Dis Child 130:1099
Drsteler MR (1975) Migrne und Vestibularapparat. J Neurol
210:253-269
Edvinsson L, Mac Kenzie ET, Mc Culloch J, Uddman R (1988)
Nerve supply and receptor mechanisms in intra- and extracerebral blood vessels. In: Oie sen J, Edvinsson L (eds) Basic
mechanisms of headache. Elsevier, Amsterdam, pp 129-144
Eeg-Olofsson 0, dkvist L, Lindskog U, Andersson B (1982)
Benign paroxysmal vertigo in childhood. Acta Otolaryngol
(Stockh) 93:283-289
Elliot MA, Peroutka SI, Welch S, May EF (1996) Familial hemiplegic migraine, nystagmus, and cerebellar atrophy. Ann Neurol
39:100-106
Estol C, Caplan LR, Pessin MS (1996) Isolated vertigo: an uncommon manifestation of vertebrobasilar ischemia. Cerebrovasc
Dis 6 (SuppI2):161
Eviatar L (1981) Vestibular testing in basilar artery migraine. Ann
NeuroI9:126-130
Eviatar L, Eviatar A (1974) Vertigo in childhood. Clin Pediatr
13:940
Fenichel GM (1967) Migraine as a cause of benign paroxysmal
vertigo in childhood. J Pediat 71 : 114-115
Ferrari MD, James MH, Bates D, Pilgrim A, Ashford E, Anderson E,
Anderson BA, Nappi G (1995) Cerebral blood fiow during
migraine attacks without aura and effect of sumatriptan. Arch
NeuroI52:135-139
Foerster 0 (1936) Sensible kortikale Felder. In: Bumke 0, Foerster
0, (eds) Handbuch der Neurologie. Springer, Berlin, pp 358-449
Gancher ST, Nutt JG (1986) Autosomal dominant episodic ataxia:
a heterogeneous syndrome. Mov Dis 1:329-353
Ganji S, Hellmann S, Stagg S, Forlow J (1993) Episodic coma due
to acute basilar artery migraine: correlation of EEG and brain
stern auditory evoked potential patterns. Clin Electroencephal
24:44-48
Goadsby PI, Lance JW (1988) Brainstem effects on intra- and

Migraine and vertigo


extracerebral circulations. In: Olesen J, Edvinsson L (eds)
Basic mechanisms of headache. Elsevier, Amsterdam, pp
413-426
Golden GS, French JH (1975) Basilar artery migraine in young
children. Pediatrics 56:722-726
Griggs RC, Nutt JG (1995) Episodic ataxias as channelopathies.
Ann NeuroI37:285-287
Harker LA (1996) Migraine associated vertigo. In: Baloh RW,
Halmagyi GM (eds) Disorders ofthe vestibular system. Oxford
University Press, Oxford, pp 407-417
Harker LA, Rassekh CH (1987) Episodic vertigo in basilar artery
migraine. Otolaryngol Head Neck Surg 96:239-250
Headache Classification Committee of the International
Headache Soeiety (1988) Classification and diagnostic criteria
for headache disorders, cranial neuralgias and faeial pain.
Cephalalgia 8 (Suppl 7) 1-93
Heinrich JB, Sandercock PAG, Warlow CP, Jones LN (1986) Stroke
and migraine in the Oxfordshire Community Stroke Project. J
Neurol 233:257-262
Heveroch J (1925) La migraine vestibulaire. Rev Neurol
33:925-929
Hinchcliffe R (1967) Headache and Meniere's disease. Acta
Otolaryngol (Stockh) 63:384-390
Hoekstra-van Dalen RAH, Cillessen JPM, Kappelle LI, van Gijn J
(1996) Cerebral infarcts assoeiated with migraine: clinical features, risk factors and follow-up. J NeuroI43:511-515
Hopf HC (1987) Vertigo and masseter paresis. A new local brain
stern syndrome probably of vascular origin. J NeuroI235:42-45
Jamada T, Dickens QS, Arensdorf K, Corbett J, Kimura J (1986)
Basilar migraine: polarity-dependent alteration of brain stern
auditory evoked potential. Neurology 36:1256-1260
Joutel A, Bousser MG, Bioussek V, Labange P, Chabriat H, Nibbio
A, Maciazek I, Meyer B, Bach MA, Weissenbach I, Lathrop GM,
Tournier-Lasserve E (1993) A gene for familial hemiplegie
migraine maps to chromosome 19. Nature Genet 5:40-45
Kayan A (1973) Migraine and related disorders of hearing and
balance. (Leading article.) Hemicrania 4:3-4
Kayan A (1984) Migraine and vertigo. In: Dix MR. Hood JD (eds)
Vertigo. Wiley, Chichester, pp 249-265
Kayan A, Hood JD (1984) Neuro-otological manifestations of
migraine. Brain 107:1123-1142
Kim HN, Kim H, Park IY, Kim GR, Chung IH (1990) Variability of
the surgical anatomy of the neurovascular complex of the cerebellopontine angle. Ann Otol Rhinol LaryngoI90:288-296
Koehler B (1980) Benign paroxysmal vertigo in childhood. A
migraine equivalent. Eur J Pediatr 134: 149
Koenigsberger MR, Chutorian AM, Gold AP, Schvey MS (1970)
Benign paroxysmal vertigo of childhood. Neurology
20:1108-1113
Kuritzky A, Toglia UJ, Thomas D (l981a) Vestibular function in
migraine. Headache 21:110-112
Kuritzky A, Ziegler DK, Hassanein R (l981b) Vertigo, motion sickness and migraine. Headache 21:227-231
Lanzi G, Ballotini U, Fazzi E, Tagliasachi M, Manfrin M, Migra E
(1994) Benign paroxysmal vertigo of childhood: a long term
follow-up. Cephalalgia 14:458-460
Lauritzen M (1987) Cerebral blood flow in migraine and cortical
spreading depression. Acta Neurol Scand (Suppll13) 76: 1-40
Lees F, Watkins SM (1963) Loss of consciousness in migraine.
Lancet ii:247
Leliever WC, Barber HO (1981) Recurrent vestibulopathy.
Laryngoscope 91:1-6
Lempert T, Menzhausen L, Tiel-Wilck K (1993) Migrne: Eine
Differentialdiagnose des episodischen Schwindels. Nervenarzt
64:121-126
Meniere P (1861) Memoire sur les lesions de l'oreille interne donnant lieu 11 des symptomes de congestion cerebrale apoplectiforme. Gaz Med (Paris) Ser 3,16:597-601

339
Mira E, Piacentino G, Lanzi G, Balottin U (1984) Benign paroxysmal vertigo in childhood. Diagnostic significance of vestibular
examination and headache provocation tests. Acta Otolaryngol
(Stockh) SuppI406:271-274
Moretti G, Manzoni GC, Caffarra P, Parma M (1980) "Benign
recurrent vertigo" and its connection with migraine. Headache
20:344-346
Moskowitz MA (1990) Basic mechanisms in vascular headache.
Neurol Clin 8:801-815
Mri RM, Meienberg 0 (1993) Drehschwindel und Migrne.
Schweiz Med Wochenschr 123: 1331-1336
Olesen J (1991) Migraine and other headaches: the vascular mechanisms. Raven Press, New York
Olsson JE (1991) Neurootologie findings in basilar migraine.
Laryngoscope 101:1-41
Pearce J (1971) Some aetiological factors in migraine. In: Cumings
JW (ed) Background to migraine. Heinemann Medical, London,
pp 1-7
Penfield W, Jasper H (1954) Epilepsy and the functional anatomy
of the human brain. Little Brown, Boston
Pryse Phillips W, Findlay H, Tugwell P, Edmeads I, Murray TJ,
Nelson RF (1992) A Canadian population survey on the clinical,
epidemiologie and soeietal impact of migraine and tensiontype headache. Can J Neurol Sei 19:333-339
Ptacek LI, Tawil R, Griggs RC, Meola G, McManis P, Barohn RJ,
Mendell JR, Harris C, Spitzer R, Santiago F, Lepper TMF (1994)
Sodium channel mutations in acetazolamide-responsive
myotonia congenita, paramyotonia congenita, and hyperkalemic periodic paralysis. Neurology 44:1500-1503
Rassekh CH, Harker LA (1992) The prevalence of migraine in
Menieres disease. Laryngoscope 102:135-138
Raskin NH (1993) Auto and prophylactic treatment of migraine:
practical approaches and pharmacological rationale. Neurology
43 (SuppI3):S39-S42
Russel MB, Olesen J (1993) The genetics of migraine without aura
and migraine with aura. Cephalalgia 13:425-428
Russel MB, Iselius L, Olesen J (1996) Migraine without aura and
migraine with aura are inherited disorders. Cephalalgia
16:305-309
Sanner G, Bergstrm B (1979) Benign paroxysmal torticollis in
infancy. Acta Pediatr Scand 68:219-223
Schiller F, Hedberg WC (1960) An appraisal of positional nystagmus,AMA.Arch NeuroI2:309-316
Selby G, Lance JW (1960) Observations on 500 cases of migraine
and allied vascular headaches. J Neurol Neurosurg Psychiatry
23:23-32
Seto H, Shimizu M, Futatsuya R, Kageyama M, Wu Y, Kamei T,
Shibata R, Kakishita M (1994) Basilar artery migraine.
Reversible ischemia demonstrated by Tc-99m HMPAO brain
SPECT. Clin Nucl Med 19:215-218
Silberstein SD, Lipton RB (1994) Overview of diagnosis and treatment of migraine. Neurology 44 (Suppl 7):S6-S16
Slater R (1979) Benign recurrent vertigo. J Neurol Neurosurg
Psychiatry 42:363-367
Snyder CH (1969) Paroxysmal torticollis in infancy. Am J Dis
Child 117:458
Stewart WF, Lipton RB, Celentano DD, Reed ML (1992) Prevalence
of migraine headache in the United States-relation to age, race,
income, and other sociodemographic factors. JAMA 267:64-69
Stewart WF, Schecter A, Lipton RB (1994) Migraine heterogeneity:
disability, pain intensity, attack frequency, and duration.
Neurology 44 (SuppI4):S24-S39
Strupp M, Brning R, Wu RH, Reiser M, Brandt Th (1998)
Diffusion-weighted MRI in transient global amnesia: elevated
signal intensity in the left medial temporal lobe in 7 of 10
patients. Ann NeuroI43:164-170
Sturzenegger MH, Meienberg 0 (1985) Basilar artery migraine: a
follow-up study of 82 cases. Headache 25:408-415

340
Symonds CP (1926) Vertigo. Postgrad Med J 1:63-66
Toglia JU, Kuritzky A, Thomas D (1981) Common migraine and
vestibular function: Electronystagmographic study and pathogenesis.Ann OtoI90:267-271
Vahedi K, Joutel A, Van Bogaert P, Ducros A, Maciazeck J, Bach JF,
Bousser MG, Tournier-Lasserve E (1995) A gene for hereditary
paroxysmal cerebellar ataxia maps to chromosome 19p. Ann
Neuro137:289-293

Vertigo
Watson P, Steele JC (1974) Paroxysmal dysequilibrium in the
migraine syndrome of childhood. Arch Otolaryngol 99: 177 -179
Weiller C, May A, Limmroth V, Jptner M, Kaube H, v Schayck R,
Coenen HH, Diener HC (1995) Brainstem activation in spontaneous human migraine attacks. Nature Med 1:658-660
Whitty CWM (1967) Migraine without headache. Lancet
ii:283-285

Hyperviscosity syndrome and vertigo

The clinical hyperviscosity syndrome is characterised by various symptoms, such as headache,


fatigue, visual disturbanees, Raynaud's phenomenon,
hearing loss, tinnitus, vertigo, or in the worst cases,
even cerebral infarctions and heart or renal failure
(Baer et al. 1985; Fahey et al. 1965; Preston et al. 1978).
Otological symptoms - hearing loss, tinnitus, and
especially vertigo - are among the most frequent
manifestations. In the three cases described by
Andrews et al. (1988), sudden episodic attacks of
vertigo and postural imbalance, which las ted from
minutes to days, were directly related to the degree
of hyperviscosity. Positional nystagmus has also
been observed in macroglobulinaemia (Keim and
Sachs 1975; see Chap.17,p. 228).

cythemia vera (Osler 1903; Calabresi and Meyer


1959; Silverstein et al. 1962) and 10-20% in patients
with Waldenstrm's macroglobulinaemia (Logothetis
et al. 1960; Fahey et al. 1965; Ronis et al. 1966).
Otological symptoms are predominantly produced
by vascular obstruction in the venules of the peripherallabyrinthine organ, which is comparable to retinal vein congestion in hyperviscosity dis orders
(Andrews et al. 1988). The involvement of the
peripherallabyrinth rather than central brainstem
structures was inferred from the combination of vertigo syndrome with directional spontaneous
nystagmus, abnormal caloric responses, and audiovestibular test results with noother neurological
symptoms (Table 21.1). If the obstruction of the
venules and capillaries with subsequent ischaemia
causes the symptoms, then peripheral vestibular
manifestations of the hyperviscosity syndrome are
due to diminished blood ftow through the vein of the
vestibular aqueduct, through which the semicircular
canals and the utricle drain. The cochlea and the saccule drain through the vein of the cochlear aqueduct
(Schuknecht 1974); its obstruction causes hearing
loss.
The particular form of positional nystagmus seen
in macroglobulinaemia (Keim and Sachs 1975) is
probably induced by a specific gravity differential
between cupula and endolymph; this is explained by
the buoyancy hypothesis (see Chap. 17, p. 285).

Aetiology and pathomechanism

Management

Hyperviscosity is defined on the basis of the relative


viscosity of serum/blood to water. Normal relative
viscosity is about 1.8; symptoms of hyperviscosity
occur at a level of 5-6, i.e. paraprotein concentrations of ca. 40 g/l. Pathological hyperviscosity of the
blood occurs in multiple myeloma, Waldenstrm's
disease, or polycythemia vera. The incidence of
vertigo is reported to be 40% in patients with poly-

Symptomatic improvement occurs after blood


hyperviscosity is reduced, regardless of the underlying cause (Andrews et al. 1988). Management
(Wilson et al. 1991) may include

Pathological hyperviscosity of the blood may be


associated with polycythemia, hypergammaglobulinaemia or Waldenstrm's macroglobulinaemia.
Depending on the degree of hyperviscosity, episodic
vertigo mayaiso occur; it is most often caused by
venous obstruction of the peripheral labyrinth.
Symptomatic improvement is seen after blood
hyperviscosity is reduced.

The clinical syndrome

341

plasmapheresis,
chemotherapy (multiple myeloma, Waldenstrm's
disease), and

342

Vertigo

32 P chemotherapy, phlebotomy (polycythemia


vera).

Table 21.1. Peripherallabyrinthine vertigo in the hyperviscosity syndrome (e.g. polycythemia vera, Waldenstrm's disease)
Clinical syndrome
Sudden episodic vertigo (duration varying from minutes to days) with
spontaneous nystagmus, postural imbalance, and abnormal caloric
responses (with no central neurological symptoms) directly related to the
degree of blood hyperviscosity.
Headache and visual or hearing disturbances mayaiso occur.
Incidence/age/sex
Otological manifestations, especially episodic vertigo, occur in up to 40%
of cases of polycythaemia vera and in up to 20% of ca ses of Waldenstrm's
macroglobulinaemia, independently of age and sex.
Pathomechanism
Increased blood viscosity causes obstruction in the peripheral draining
vein of the vestibular aqueduct and associated capillaries leading to
hypoxia of the vestibular labyrinth.
Course/prognosis
Spontaneous changes occur with fluctuating blood viscosity and there is
symptomatic improvement with reduced hyperviscosity, regardless of the
underlying cause.
Management
Therapeutic reduction of blood hyperviscosity (plasmapheresis,
phlebotomy) is generally effective; if not, the symptoms are caused by
haemorrhage within the labyrinthine organ.
Differential diagnosis
All kinds of peripherallabyrinthine dysfunction, for example, vestibular
neuritis, Meniere's disease, basilar migraine, vestibular paroxysmia,
perilymph fistula, familial episodic ataxia.

References
Andrews JC, Hoover LA, Lee RS, Honrubia V (1988) Vertigo in the
hyperviscosity syndrome. Otolaryngol Head Neck Surg
98:144-149
Baer MR, Stein RS, Dessypris EN (1985) Chronic Iymphocytic
leukemia with hyperleukocytosis: the hyperviscosity syndrome. Cancer 56:2865-2869
Calabresi P, Meyer 0 (1959) Polycythemia vera. 1. Clinical and laboratory manifestations. Ann Intern Med 50: 1182-1202
Fahey JL, Barth WF, Solomon A (1965) Serum hyperviscosity syndrome. JAMA 192:464-467
Keim RJ, Sachs GB (1975) Positional nystagmus in assoeiation
with macroglobulinemia. Ann OtoI84:223-227
Logothetis J, Silverstein P, Coe J (1960) Neurological aspects of
Waldenstrm's macroglobulinemia. Arch Neurol 3:564-573
Osler W (1903) Chronic cyanosis with polycythaemia and
enlarged spleen: a new clinical entity. Am J Med Sei
126:187-201
Preston FE, Sokol RJ, Lilleyman JS, Winfield DA, Blackburn EK
(1978) Cellular hyperviscosity as a cause of neurological symptoms in leukemia. Br Med J i:476-478
Ronis ML, Rojer CL, Ronis BJ (1966) Otologie manifestations of
Waldenstrm's macroglobulinemia. Laryngoscope 76:513-523
Schuknecht HF (1974) Pathology of the ear. (2nd edn 1993)
Harvard University Press, Cambridge, Mass
Silverstein A, Gilbert H, Wasserman LR (1962) Neurologie complications of polycythemia. Ann Intern Med 57:909-915
Wilson JD, Braunwald E, Isselbacher KH, Petersdorf RG, Martin
JB, Fanei AS, Root RK (1991) Harrison's Prineiples of Internal
Medicine, 12th edn. McGraw-Hili, New York

SECTION F
Traumatic vertigo

Traumatic vertigo is among the most frequent


sequelae associated with head and neck injuries
(Table EI; Chap. 22) and barotrauma (Table 23.1).
Central vestibular vertigo syndromes (p. 167) due
to brainstem lesions (concussion, haemorrhage)
and typical benign paroxysmal positioning vertigo (p. 251) are weil recognised. The peripheral
end-organs and vestibular nerves mayaiso be
affected by temporal bone fractures, perilymph
fistulas, haemorrhages, or concussion of the
endolymphatic and perilymphatic spaces.
Evidence is presented for a typical traumatic
otolith vertigo (p. 349) secondary to dislodged
otoconia. The associated anxiety and secondary
gain factors in some patients doud the dear-cut
distinction between organic and psychogenic
mechanisms. Prolonged symptoms without
objective signs oE dysfunction are frequently
psychogenic.
Vertigo secondary to barotrauma (diving, pressure chamber, high-altitude exposure) is mostly
of peripheral labyrinthine origin (alternobaric
vertigo; round or oval window fistulas), or part of
decompression sickness (p. 352). When a person
with perforated tympanie membrane swims or
dives, he may experience a dangerous caloric
response produced by cold water irrigation of the
semicircular canals of the affected ear.
Vestibular dysfunction may be the inadvertent
iatrogenic consequence of medicinal, physical or
surgical procedures (Chap. 24).

Table F.t. Vertigo following head and/or neck injury


Si te

Syndrome

Mechanism

Labyrinth

Otolith vertigo

Loosening of
otoconia
Canalolithiasis

Benign paroxysmal
positioning vertigo
Loss of labyrinlhine
function
Perilymph fistula

Temporal bone
fractures,
con(ussion or
hemorrhage
Round or oval
window rupture,
temporal bone
fracture or
haemorrhag e

Vestibular nerve

Loss 01 vestibular lunction

Concussion or
haemorrhage

Brainstem or
vestibulocerebellum

All (entral vestibular vertigo


syndromes (downbeatl
upbeat nystagmus.
ocular tilt
reaction, central positional
nystagmus/vertigo. etc)

Concussion,
haemorrhage,
or stroke (e.g.
vertebral
artery dissection)

Cervical

Whiplash vertigo (not a


(Iear entity)

Not known (vascular


compression?
neuromuscular?
neurovascular?)

Cervical vertigo?
Psychogenic

345

Phobie postural vertigo

Secondary gain factor5

Anxious introspection
causes dissociation
between
efference and
efference copy

Head and neck injury

Post-traumatic dizziness or vertigo is one of the


major complaints following head (Davies and Luxon
1995; Luxon 1996) or whiplash injuries (Oosterveld
et al. 1991), not only in adults but also in children
(Fried 1980; Vartainen et al. 1985). It accounts for
increasing numbers of medicolegal claims.
There are several types of post-traumatic vertigo
(Table F. 1), and the pathogeneses are controversial
(Friedman et al. 1945; Tuohimaa 1978). What
Friedman et al. wrote about this subject in 1945 still
seems to be valid: "In a discussion of dizziness per se
following trauma, or as apart of the post-traumatic
syndrome, there are many opinions as to what are
the important factors in the evaluation of dizziness.
In general there are three schools of thought - one
group who consider dizziness, as well as the other
symptoms of the post-traumatic syndrome, to be of
psychogenic origin; another group who report it to
be exclusively of physiogenic origin; and others who
emphasise the importance ofboth types offactors".
The frequency of dizziness and disequilibrium
following head trauma may be as low as 14% for
non-hospitalised patients (Coonley-Hoganson et al.
1984), but in most studies the incidence lies somewhere around 40-60% (Friedman et al. 1945;
Gannon et al. 1978). A considerable proportion of
these dizzy subjects do not return to their preaccident work, or to equivalent work, within 5 years
of the injury (Berman and Fredrickson 1978; Eide
and Tysnes 1992). Even in cases of mild head
trauma, dizziness was reported to occur within 1
week in 53% (Levine et al. 1987) and to persist at
least 2 years in 18% (Cartlidge 1978). These statistics
are, however, of little clinical significance, and retrospective statistical analyses of abnormal oculographic findings such as spontaneous or positional
nystagmus (Lange and Kornhuber 1962; Tuohimaa
1978) or post-traumatic squint and-diplopia (Lepore
1995; Fowler et al. 1996) are of equally limited value.
Traumatic vertigo is not a clinical entity. The term
signifies only a common aetiology for a heterogeneous collection of peripherallabyrinthine and central vestibular disorders (Table EI).

It is possible to clearly distinguish several typical


forms of post-traumatic vertigo, which make up the
majority of cases. These are

347

Post-traumatic benign paroxysmal positioning


vertigo (p. 264). This is the most frequent cause
of vertigo following blunt head trauma (Davies
and Luxon 1995) or whiplash injury (Brandt and
Daroff 1980). It may manifest only days but also
weeks after head injury (Saito et al. 1986) and
may be bilateral. The late onset of symptomatology can be explained by
1. slow degeneration of the otolith organ after
labyrinth concussion,
2. the settling of dislodged otoconia in the utricular cavity before entering the semicircular
canal,or
3. the time needed for several pieces of otoconia
to form a clot (canalolith) to become causative.
Post-traumatic otolith vertigo (p. 349) without
canalolithiasis. Here labyrinthine concussion
causes loosening and dispersion of otoconia
from the macular bed, which results in une qual
heavy loads of right and left otoliths (see
traumatic otolith vertigo, p. 349).
Labyrinthine and/or eighth nerve dysfunction
with typical fra ctu res of the temporal bone. These
fractures are of two kinds: a longitudinal fracture (Fig. 9.25), which involves the middle ear
structures with dislocation of the ossicular chain
but usually spares the labyrinth and the eighth
nerve, and a transverse fracture (perpendicular
to the axis of the petrous bone), which preferably
involves the bony labyrinth or the seventh and
eighth cranial nerves in the internal acoustic
meatus, depending on whether the fracture is
located more laterally or medially. Longitudinal
fractures account for ab out 80% (temporoparietal impact); transverse fractures, for about 20%
(occipital impact) of all temporal bone fractures.
The facial and vestibulocochlear nerves are
affected in ab out 50% of transverse fractures.
The vestibular system seems less vulnerable than

Vertigo

348

the auditory system and fractures here have a


greater chance of recovery (Wennmo and
Spandow 1993). Partially preserved vestibular
function with total hearing loss is reported more
frequently than the opposite combination
(Wennmo and Svensson 1989). Acute bilateral
vestibulocochlear dysfunction has also been
described following occipital fracture (Feneley
and Murthy 1994).
Labyrinthine concussion after blunt head trauma
without fracture. Labyrinthine concussion with
vestibular dysfunction often associated with
bilateral high-tone sensorineural hearing loss is
common (Davies and Luxon 1995) and has been
ascribed to microscopic haemorrhages in the
cochlea and labyrinth. Traumatic damage to the
labyrinth can be visualised by contrast enhancement and hyperintense signals (intralabyrinthine small bleedings) in T/T 2 -weighted MR
images (Goodwin 1989). Experimental head
injury in guinea pigs showed disarrangement of
the vestibular system with lithic, exfoliate, vacuolisation of the sensory epithelia, and exfoliated
membranes in the utricular and saccular maculae with otolith separation from the saccular
maculae (Zhou et al. 1994).
Non-explosive blast injury of the ear. A slap or a
blow to the ear during sport accidents (e.g. ball
games, water sports) can seal the external
acoustic meatus and result in a sudden increase
in air pressure within the ear canal which strikes
against the tympanic membrane (Berger et al.
1994). Common symptoms are similar to those
caused by explosive injury (p. 352) and include
hearing loss, tinnitus, vertigo, disequilibrium,
earache and otorrhea.
Perilymphatie fistula following even mild blunt
head trauma (Grimm et al. 1989; Glasscock et al.
1992) or transverse temporal bone fractures
(Lyos et al. 1995).
Delayed endolymphatie hydrops (p. 88) following
labyrinthine concussion (Nadol et al. 1975) or
temporal bone fractures.
Secondary phobie postural vertigo following
recovery of organic post-traumatic vestibular
dysfunction (e.g. traumatic otolith vertigo,
benign paroxysmal positioning vertigo,
labyrinthine concussion). This should be suspected when dizziness caused by head trauma
lasts longer than 4-6 weeks without a noticeable
improvement and with normal otoneurological
test results (Huppert et al. 1994; Brandt 1996).
Vertigo and/or Wallenberg's syndrome due to
vertebral artery disseetion. This has been repeatedly described following neck injury or chiro-

practic neck manipulations (Vibert et al. 1993;


Shimizu et al. 1992).
Central vestibular syndromes due to concussion
of vestibular nuclei or eentral vestibular pathways, particularly in the caudal brainstem.

Trauma may simultaneously or separately damage


peripheral and central vestibular structures. Up to
50% of patients with mild head injury (Williams et
al. 1990) develop post -concussional syndrome
(Fisher 1966; Rutherford et al. 1990). The duration of
unconsciousness or post-traumatic amnesia is the
usual indicator by which the severity of a closed
head injury without focal neurological deficits may
be evaluated. The post-traumatie syndrome (dizziness, headache, tinnitus, hearing loss, blurred vision,
diplopia, anxious irritability, depression, emotional
lability, a decrement in attentional and information
processing and fatigue) is attributed to the diffuse
microscopic changes that accompany mild concussion, the most common form of closed head injury.
Even mild concussions can cause significant attentional and information processing impairments lasting weeks or months, in the absence of any apparent
neurological problems (Hugenholtz et al. 1988).
Animal experiments have shown that even minor
head injuries can produce petechial cerebral haemorrhages due to distortion forces, especially within
the brainstem and the vestibular nuclei (DennyBrown and Russel 1941; Nakamura 1967; Jellinger
1967).

Finally, high-impact aerobics have been accused


of injuring the delicate otoliths and cochlea, causing
a syndrome of vertigo, tinnitus, balance dysfunction
and hearing loss (Weintraub 1994).
Dizziness is also one of the most common distressing symptoms resulting from so-called
whiplash injuries (Compere 1968; Rubin 1973;
Hinoki 1985). A large proportion of these cases may
be due to otolithic vertigo, with or without
canalolithiasis. Other, older explanations are based
on a number of obscure hypothetical pathogeneses
such as compression of the proximal vertebral artery
against the transverse process of the seventh cervical
vertebra by contracted cervical fascia (Compere
1968), a neuromuscular mechanism (Gray 1956), a
neurovascular mechanism (Hyslop 1952; Weeks and
Travelli 1955), or an "overexcitation of the cervical
and lumb ar proprioceptors" (Hinoki 1985).
Whiplash vertigo also raises the question ofhow cervical vertigo may be diagnosed (p. 441). It is interesting that retrospective studies on traumatic dizziness
following head or whiplash injuries have also indicated that the diagnosis seems to depend on whether
the investigators are otolaryngologists or neurologists.
A neurological study on 100 patients with post-

349

Head and neck injury

traumatic dizziness found that only eight had central


vestibular abnormalities (Davies and Luxon 1995),
whereas an ENT study on 318 patients suffering
from dizziness after a whiplash trauma reported
central-vestibular dysfunction in 258 (Kortschot and
Oosterveld 1994). The latter is not supported by our
own experience.
Chronic post-traumatic dizziness or disequilibrium
which persists for months or years without abnormal otoneurological or neuro-ophthalmological
findings is most likely psychogenic, especially if
accompanied by chronic headache (tension or cervicogenic) and depression. All peripheral, and most
central, vestibular syndromes gradually improve
with time, due to functional recovery or central compensation (p. 55). Careful assessment of vestibular
function is needed before establishing a treatment
regimen. This may consist of exercises designed to
readjust the vestibular responses to enhance central
compensation or sensory substitution of the vestibular deficit (Herdman 1990).

Traumatic otolith vertigo


Patients often describe their post-traumatic vertigo
as a non-rotatory to-and-fro vertigo which is particularly associated with head acceleration and an
unsteadiness of gait similar to walking on pillows.
Since these post-traumatic symptoms resemble
otolith dysfunction in many patients, one can speculate that this vulnerable accelerometer is affected by
trauma (Brandt and Daroff 1980). The calcareous
material embedded in the gelatinous matrix may
loosen, resulting in unequal loads on the macula
beds and a tonus imbalance between the two. This
has been shown in centrifuging experiments in
animals (Hasegawa 1933; Igarashi and Nagaba 1968).
Engineering accelerometers are just as vulnerable.
Such a mechanism also fits the finding of DeWit and
Bles (1975) that postural sway is increased after
headshaking in dizzy patients who have suffered a
concussion and depends more on visual stabilisation
than in normal subjects. The authors erroneously
believed their findings resulted from brainstem conCUSSlOn.

A considerable proportion of traumatic vertigo


can be assumed to be due to dislodged otoconia,
with or without concurrent benign paroxysmal positional vertigo, which depends on the position of the
debris within the membraneous labyrinth. No clinical test is yet available to establish a diagnosis of
traumatic otolithic vertigo, but our own unpublished
measurements of transient deviation of the subjec-

tive visual vertical in these patients provide evidence


of the condition.
Central compensation (rearrangement) would
account for the gradual recovery within days or
weeks, thus supporting the view that exercise is the
best therapy.

References
Berger G, Finkelstein Y, Harell M (1994) Non-explosive blast
injury of the ear. J Laryngol Otol108:395-398
Berman JM, Fredrickson (1978) Vertigo after head injury: a five
year follow-up. J OtolaryngoI7:237-245
Brandt Th (1996) Phobic postural vertigo. Neurology
46: 1515-1519
Brandt Th, Daroff RB (1980) The multisensory physiological and
pathological vertigo syndromes.Ann NeuroI7:195-203
Cartlidge NEF (1978) Postconcussional syndrome. Scott Med J
23:103
Compere WE (1968) Electronystagmographic findings in patients
with "whiplash" injuries. Laryngoscope 78: 1226-1233
Coonley-Hoganson R, Sachs N, Desai BT, Whitman S (1984)
Sequelae associated with head injuries in patients who were not
hospitalized: a follow-up survey. Neurosurgery 14:315-317
Davies RA, Luxon LM (1995) Dizziness following head injury: a
neuro-otological study. J NeuroI242:222-230
Denny-Brown R, Russel WR (1941) Experimental cerebral concussi on of the brain. Laryngoscope 50:921
DeWit G, Bles W (1975) A stabilographic study of the role of optic
stimuli in maintaining the postural position in patients suffering from postconcussional dizziness. Agressologie 16D:9-14
Eide PK, Tysnes O-B (1992) Early and late outcome in head injury
patients with radiological evidence of brain damage. Acta
Neurol Scand 86:194-198
Feneley MR, Murthy P (1994) Acute bilateral vestibulo-cochlear
dysfunction following occipital fracture. J Laryngol Otol
108:54-56
Fisher CM (1966) Concussion amnesia. Neurology 16:826-830
Fowler MS, Wade DT, Richardson AJ, Stein JF (1996) Squints and
diplopia seen after brain damage. J NeuroI243:86-90
Fried MP (1980) The evaluation of dizziness in children.
Laryngoscope 90: 1548-1560
Friedman AP, Brenner C, Denny-Brown D (I945) Post-traumatic
vertigo and dizziness. J Neurosurg 21:36-46
Gannon P, Wilson GN, Roberts ME, Pearse HJ (1978) Auditory and
vestibular damage in head injuries at work. Arch Otolaryngol
104:404-408
Glasscock ME, Hart MI, Rosdeutscher JD, Bhansali SA (1992)
Traumatic perilymphatic fistula: how long can symptoms persist? A follow-up report. Am J OtoI13:333-338
Goodwin WJ (1989) Temporal bone fractures. Radiol Clin North
Am 16:651-659
Gray L (1956) Extra-labyrinthine vertigo due to cervical muscle
lesions. J Laryngol Otol 70:352-361
Grimm RJ, Hemenway WG, LeBray PR, Black FO (1989)The perilymph fistula syndrome defined in mild head trauma. Acta
Otolaryngol (Suppl) 464:1-40
Hasegawa T (1933) Die Vernderungen der labyrinth ren Reflexe
bei zentrifugierten Meerschweinchen. Pflgers Arch Ges
PhysioI232:454-465
Herdman SJ (1990) Treatment of vestibular dis orders in
traumatically brain-injured patients. J Head Trauma Rehabil
5:63-76

350
Hinoki M (1985) Vertigo due to whiplash injury: a neuro-otological
approach. Acta Otolaryngol (Stockh) SuppI419:9-29
Hugenholtz H, Stuss DT, Stethem LL, Richard MT (1988) How
long does it take to recover from a mild concussion?
Neurosurgery 22:853-858
Huppert D, Kunihiro T, Brandt Th (1994) Phobic postural vertigo
(154 patients): its association with vestibular disorders. J Audiol
Med 4:97-103
Hyslop G (1952) Intra-cranial circulatory complication of injuries
to the neck. Bull NY Acad Med 28:729-733
Igarashi M, Nagaba M (1968) Vestibular end-organ damage in
squirrel monkeys after exposure to intensive linear acceleration. In: Third symposium on the role of the vestibular organs
in space exploration, NASA SP-152, pp 63-67
Jellinger K (1967) Hufigkeit und Pathogenese zentraler
Hirnlsionen nach stumpfer Gewalteinwirkung auf den
Schdel. Wien Z Nervenheilk 25:223
Kortschot HW, Oosterveld WJ (1994) Otoneurologic disorders
after cervical whiplash trauma. Orthopde 23:275-277
Lange G, Kornhuber HH (1962) Zur Bedeutung peripher- and
zentral-vestibulrer Strungen nach Kopftraumen. Arch
Ohren-Heilk u Z Hals-Heilk 179:366-385
Lepore FE (1995) Disorders of ocular motility following head
trauma. Arch Neurol 52:924-926
Levine HS, Mattis S, Ruff RM, Eisenberg HM, Marshall LF, High
WM jr, Frankowski RF (1987) Neurobehavioral outcome following minor head injury: a three-center study. J Neurosurg
66:234-243
Luxon LM (1996) Posttraumatic vertigo. In: Baloh RW, Halmagyi
GM (eds) Disorders of the vestibular system. Oxford University
Press, Oxford, pp 381-395
Lyos AT, Marsch MA, Jenkins HA, Coker NJ (1995) Progressive
hearing loss after transverse temporal bone fracture. Arch
Otolaryngol Head Neck Surg 121:795-799
Nadol JB, Weiss AD, Parker SW (1975) Vertigo of delayed onset
after sudden deafness.Ann Otol Rhinol LaryngoI84:841-846
Nakamura N (1967) Mechanism of head trauma. Clin Med 9: 1131

Vertigo
Oosterveld WJ, Kortschot HW, Kingma GG, de Jong HAA, Saatci
MR (1991) Electronystagmographic findings following cervical
whiplash injuries. Acta Otolaryngol (Stockh) 111:201-205
Rubin W (1973) Whiplash with vestibular involvement. Arch
OtolaryngoI97:85-87
Rutherford WH, Merret JD, McDonald JR (1990) Sequelae of concussion caused by minor head injuries. Lancet i: 1-4
Saito Y, Ishikawa T, Makiyama Y, Hasegawa M, Shigihara S,
Yasukata J, Ishiyama E, Tomita H (1986) Neuro-otological study
of positional vertigo caused by head injury. Auris Nasus Larynx
(Tokyo) 13:S69-S73
Shimizu J, Nakagawa Y, Fuji Y, Nakase H, Mannen T (1992)
Wallenberg's syndrome due to vertebral artery dissection following minimal neck injury: report of two cases. Rinsho
Shinkeigaku 32:430-435
Tuohimaa P (1978) Vestibular disturbances after acute mild head
injury. Acta Otolaryngol (Stockh) SuppI359:1-67
Vartianien E, Karjalainen S, Karja J (1985) Vestibular disorders
following head injury in children. Int J Pediatr 9: 135-141
Vibert D, Rohr Le Floch J, Gauthier G (1993) Vertigo as manifestation of vertebral artery dissection after chiropractic neck
manipulations. ORL J Otorhinolaryngol Relat Spec 55:140-142
Weeks V, Travelli J (1955) Postural vertigo due to trigger areas in
sterno-cleidomastoid muscle. J Pediat 47:315-327
Weintraub MI (1994) Vestibulopathy induced by high impact aerobics. A new syndrome: discussion of 30 cases. J Sports Med
Phys Fitness 34:56-63
Wennmo C, Svensson C (1989) Temporal bone fractures. Acta
Otolaryngol Suppl (Stockh) 468:379-383
Wennmo C, Spandow 0 (1993) Fractures of the temporal bonechain incongruencies. Am J Otolaryngol 14:38-42
Williams DH, Levin HS, Eisenberg HM (1990) Mild head injury
classification. Neurosurgery 27:422-428
Zhou D, Xu W, He L (1994) Histopathology of nonacoustic
labyrinth following head injury in guinea pigs. Chung-HuaErh-Pi-Yen-Hou-Ko-Tsa-Chih 29:350-352

Vertigo due to barotrauma

Ihis type of vertigo is associated with exposure to


alterations in ambient pressure, either an increase
(diving, pressure chamber, explosions) or a decrease
(fIying, altitude chambers). Ihe atmosphere exerts
an absolute pressure of 760 mmHg (lOB mbar) at
sea level, the standard one atmosphere absolute (1
AIA) pressure. Changes of pressure in water increase
linearly with increasing depth: one atmosphere is
added for each 10 m. Ihis increased pressure is balanced by breathing air delivered at the new ambient
pressure and by equalizing the press ure in all gascontaining body cavities to ambient (Farmer and
Ihomas 1976; Margulies 1987). Ihe volume of gas
varies inversely with ambient pressure. It is this pressure-volume relationship that mostly causes barotrauma. Ihe likelihood of damage to the Eustachian
tube and middle and inner ear increases as the rate
of change of external pressure increases, because
large pressure differentials are produced in these
areas.
Fistulas of the round and oval windows are potential hazards when exposed to hyperbaric conditions
during diving (Goodhill1971). Ihis danger has been
demonstrated in animal experiments (guinea-pigs)
using apressure chamber (Nakashima et al. 1988).
Cochleovestibular damage occurs more frequently,
however, during decompression (see decompression
sickness, p. 352) than during compression under
hyperbaric conditions. About 1% of divers exhibit
some kind of barotrauma during decompression; 5%
of these cases involve hearing or balance (Head
1984). Ihe time span between barotrauma and onset
of audiovestibular dysfunction may help to distinguish decompression sickness from perilymphatic
fistulas and to determine the appropriate therapy.
Individuals who experience a 10ss of hearing, vertigo, and nausea within minutes to a few hours after a
decompression dive or high-altitude exposure
should be suspected of having decompression siekness and should be recompressed. If symptoms do
not disappear or if audiovestibular dysfunction
occurs within 3 days after a possible barotrauma,

then round or oval window fistula should be suspected (Pullen 1992). Iable 23.1 lists the causes of
vestibular barotrauma.
Table 23.1. Vestibular barotrauma
Alternobarie vertigo
Blast injury to the ear
Decompression sickness
Oval window rupture (luxated stapes)
Round window rup!ure
Basilar membrane rupture

Alternobarievertigo
Alternobaric vertigo is a poorly understood condition of transient rotatory or to -and-fro vertigo due
to pressure changes in the middle ear. It primarily
affects about 10 - 25% of aircrews and divers
(Armstrong and Heim 1937; Lundgren and Malm
1966; Luxon 1996). Ihe on set of vertigo and nystagmus is preceded bya sensation of fullness of the ear.
Ihis condition may last for seconds to hours and
resolve spontaneously. In divers it occurs more fre quently on ascent than descent. In aircrews it occurs
during ascent of the aircraft. Frequently, susceptible
individuals complain of upper respiratory tract congestion, which presumably affects Eustachian tube
patency. In such cases, vertigo is usually relieved by
equilibration of middle ear atmospheric pressure
(Luxon 1996).Alternobaric vertigo may reflect inadequate stimulation of semicircular canals or otoliths.
It is thought to be caused by an inability to equalise
one or both ears during a sudden middle ear overpressurisation (poor Eustachian tube function) and
rapidly shifted positions of the round and oval
(movement of stapes footplate) windows (Margulies
1987). Ihe incidence of alternobaric vertigo in pilots
increased as aircrafts became more powerful.
Pressure-breathing suits have been designed to
enhance pilots' tolerance to alterations in gravity

351

Vertigo

352

(Wicks 1989). Patients with perilymph fistulas may


become symptomatic when exposed to a sudden
change in ambient pressure, for example, when travelling through mountains (p. 99).

Blast injury
Blast injuries to the ear are generally due to situations of violence or industrial accidents (Chait et al.
1989). Shock waves following explosions are characterised by an initial, short positive pressure (condensation phase) followed by a longer negative pressure
(rarefaction phase) Oahrsdoerfer 1979). Damage to
the ear most often includes tympanic membrane
perforation, sensorineural hearing loss, and sometimes reversible vestibular dysfunction, which is
often overlooked (Shupak et al. 1993). Utricular and
saccular ruptures have been found histopathologically in persons who died from blast trauma (Kerr
and Byrne 1975a). Benign paroxysmal positioning
vertigo, which rarely follows blast injuries (Kerr and
Byrne 1975b; Chait et al. 1989), can be explained by
traumatic loosening of otoconia with secondary
canalolithiasis (p. 251). Blast injury mayaiso cause
perilymph fistulas (p. 99), with somatosensory hearing loss and episodic vertigo.

Decompression sickness
Decompression sickness most frequently occurs
during scuba diving (Head 1984; Arness 1997) or
high-altitude exposure, either in an aircraft or an
altitude chamber (Black and DeHart 1992; Rudge
1992). Exceptional persons such as astronauts who
perform extravehicular activities (space walk) are
also at risk (Conklin et al. 1996). Predisposing factors for the development of decompression sickness
include exertion, injury, cold, dehydration, alcohol
consumption, repetitive diving, flight after diving,
age, and obesity (Arness 1997). There is a trend
toward greater susceptibility with increasing age,
particularly for individuals over 42 years of age
(Sulaiman et al. 1997). The relative risk for females
has been reported to be 3-4 times that of males
(Weien and Baumgartner 1990).
During decompression the nitrogen dissolved in
the blood and tissues is released and may form
bubbles in blood vessels or tissue. In vitro techniques measuring bubble growth at various altitudes

have shown that the growth rates are very dependent


on the altitude. For example, bubbles do not grow
beyond a small size at altitudes below 5400 m or
when the fluid surrounding the bubbles is cleared of
supersaturated gases (Olson and Krutz 1991). These
findings correspond to the clinical experience that
decompression sickness symptoms seldom develop
below 5400 m, and require some time to occur after
arrival at high altitudes, rarely beginning less than 4
hours later (Kumar et al. 1990; Olson and Krutz
1991).
Bubbles are the cause of the well-known decompression sickness, whose characteristic joint pain
(bends) arises from bubble formation in periarticular sites. Neurological effects of decompression sickness may be caused by either bubbles in the nervous
tissue or bubble emboli. They may affect cerebral tissue or the sensory end-organs, producing temporary
or permanent vertigo, tinnitus, and sensorineural
hearing loss. Involvement of labyrinthine end-arte ries
results in temporary or permanent cochleovestibular
symptoms. It has also been suggested that changes in
microcirculation causing local capillary stasis,
osmotic changes, and oedema of the vessel walls
contribute to labyrinthine dysfunction in the hyperbaric state and during decompression. According to
Edmonds and Freeman (1972), damage may occur
on rare occasions during the changeover from oxyhelium to compressed air as arespiratory gas before
decompression starts, when bubbles form at the
interface of the middle and the inner ear (due to
counter-diffusion at the round window). A porcine
model of neurological decompression sickness and
its treatment (Broome and Dick 1996) allows the
prospective evaluation of risk reduction and treatment strategies.

Management
Decompression sickness is managed by early recompression (hyperbaric oxygen) in a compression
chamber. This is the most effective treatment regardless of whether it is caused by scuba diving (Arness
1997) or high-altitude exposure (Weien and
Baumgartner 1990). Early recognition and recompression are essential for resolution without sequelae,
but often the treatment is still beneficial if performed days or even weeks after barotrauma. The
use of heparin can be recommended for secondary
thrombosis (Margulies 1987). Vestibular symptoms
due to decompression bubble formation must be differentiated clinically from barotrauma fistula, which
sometimes requires diagnostic tympanotomy.

Vertigo due to barotrauma

Round and oval window fistula


caused by barotrauma
The middle ear cavity normally contains air at
atmospheric pressure, which is obtained by air
exchange from the nasopharynx via the Eustachian
tube, whose orifice opens actively when swallowing
or yawning. Since ambient pressure changes are
transferred straight through the external auditory
meatus toward the tympanie membrane and middle
ear, failure of ventilation by the Eustachian tube
causes painful pressure gradients. Relative overpressurisation of either the inner or the middle ear
may occur with tube dysfunction or with rapid pressure changes during physical exertion or forced
Valsalva manoeuvre. This may result in a rupture of
the tympanie membrane or the formation of a fistula
of the round or oval windows. Patients may present
with fluctuating progressive sensorineural hearing
loss, episodic vertigo, disequilibrium, tinnitus and
aural pressure, a clinical picture that mimics that of
Meniere's disease (Luxon 1996). A history of barotrauma (e.g. physical exertion, Valsalva manoeuvre,
flying, diving) is then the key to the correct diagnosis.
Extreme increases in pressure in the middle ear
can, on rare occasions, luxate the stapes medially
toward the inner ear. This may occur when playing a
brass instrument if the Eustachian tube is pathologically wide and therefore open. If a perilymph
leak occurs as a result of the luxation, the condition

353

is known as fistula of the oval window (see perilymph fistula, p. 102). Symptoms are sensorineural
hearing loss and episodic vertigo, either of the
semicircular canal type (p. 28) or the otolith type
(p. 29), because utricle and saccule are adjacent to
the stapes footplate. Luxation of the stapes footplate
without continuous perilymph leakage - due to
barotrauma - can manifest as an otolith Tullio
phenomenon (Brandt et al. 1988; Ildiz and Dunbar
1994), with sound-induced paroxysmal eye-head tilt,
oscillopsia, and postural imbalance (see otolith
Tullio phenomenon, p. 107). The most probable
mechanism is mechanical stimulation of the otoliths
by the luxated stapes footplate, which is driven to
pathologically large amplitudes by the soundinduced stapedius reflex (Brandt et al. 1988).
Possible pathways of implosive and explosive forces
leading to cochleovestibular damage are illustrated
in Fig. 23.l.
The stapes footplate may be luxated toward the
middle ear, or the thin round window membrane
may rupture when rapid changes of the intracranial
pressure are transmitted to the membranous
labyrinth, as for example in persisting infantile
cochlear aqueduct (p. 377). In this case, increases in
cerebrospinal fluid press ure during physical exertion, when lifting weights for example, raise perilymph pressure within the scala tympani. The
basilar membrane mayaiso be ruptured by the same
mechanism, resulting in a combination of sudden
sensorineural hearing loss and vertigo. This is of
particular concern to divers during des cent, because

Focial_ve

....... 1IHt_ nerve

CocNear_
MickIe ... .... - ... .
lIHtibule . .... . .. .. ... .

Eustachian lube ............ .

. ......... l!Muno<'s _ _
..... Sc&1a....... .. BaiIat lYW11IIrAne
..... . Sc&Ia~

Fig.23.1.

Possible pathways of implosive and explosive forces leading to cochleovestibular damage. (From Goodhill 1971.)

354

they are subject to an increased cerebrospinal fluid


pressure, heightened by physical exertion, whereas
middle ear pressure is low compared with ambient
pressure (Head 1984). Goodhill (1971) has described
the "implosive" and "explosive" pathways of pressure
forces (Fig. 23.1) responsible for the occurrence of
barotraumatic oval or round window fistulas (for
management, see perilymphatic fistulas, p. 105).
Finally, persons with tympanie membrane perforation or one external meatus blocked by wax may
experience a typical and dangerous caloric response
due to cold water irrigation of the horizontal semicircular canal when swimming or diving.

References
Armstrong HG, Heim IW (1937) The effect of fiight on the middle
ear.JAMA 109:417-421
Arness MK (1997) Scuba decompression illness and diving fatalities
in an overseas military community. Aviat Space Environ Med
68:325-333
Black WR, DeHart RL (1992) Decompression sickness: an increasing risk far the private pilot. Aviat Space Environ Med
63:200-202
Brandt Th, Dieterich M, Fries W (1988) Otolithic Tullio phenomenon typically presents as paroxysmal ocular tilt reaction. Adv
Oto Rhino LaryngoI42:153-156
Broome JR, Dick EJ (1996) Neurological decompression illness in
swine. Aviat Space Environ Med 67:207-213
Chait RH, Casler J, Zajtchuk JT (1989) Blast injury of the ear: historical review.Ann Otol Rhinol Laryngol98 (SuppI140):9-12
Conklin J, Kumar KV, Powell MR, Foster PP, Waligora JM (1996) A
probabilistic model of hypobaric decompression sickness
based on 66 chamber tests. Aviat Space Environ Med
67:176-183
Edmonds C, Freeman PB (1972) Inner ear barotrauma. Arch
OtolaryngoI95:556-563
Farmer IC, Thomas WG (1976) Ear and sinus problems in diving.
In: Strauss (ed) Diving medicine. Grune and Stratton, New
York, pp 109-133

Vertigo
Goodhill V (1971) Sudden deafness and round window rupture.
Laryngoscope 81:1462-1474
Head PW (1984) Vertigo and barotrauma. In: Dix MR, Hood JD
(eds) Vertigo. Wiley, Chichester, pp 199-215
Ildiz F, Dunbar A (1994) A case of Tullio phenomenon in a subject
with oval window fistula due to barotrauma. Aviat Space
Environ Med 65:67-69
Jahrsdoerfer R (1979) The effects of impulse noise on the eardrum
and middle ear. Otolaryngol Clin North Am 12:515-520
Kerr AG, Bryne JET (1975a) Concussive effects ofbomb blast on
the ear. J Laryngol Otol 89: 131-143
Kerr AG, Bryne JET (1975b) Blast injuries to the ear. BMJ
1:559-561
Kumar KV, Waligora JM, Calkins DS (1990) Threshold altitude
resulting in decompression sickness. Aviat Space Environ Med
61:685-689
Lundgren CEG, Malm LU (1966) Alternobaric vertigo among
pilots.Aerospace Med 37:178-180
Luxon LM (1996) Post-traumatic vertigo. In: Baloh RW, Halmagyi
M (eds) Disorders of the vestibular system. Oxfard University
Press, Oxford pp 381-395
Margulies ADC (1987) A short course in diving medicine. Ann
Emerg Med 16: 689-701
Nakashima T, ltoh M, Sato M, Watanabe Y, Yanagita N (1988)
Auditory and vestibular dis orders due to barotrauma. Ann Otol
Rhinol LaryngoI97:146-152
Olson RM, Krutz RW (1991) Significance of delayed symptom
onset and bubble growth in altitude decompression sickness.
Aviat Space Environ Med 62:296-299
Pullen FW (1992) Perilymphatic fistula induced by barotrauma.
Am J Otol13:270-272
Rudge FW (1992) Altitude-induced arte rial gas embolism: a case
report. Aviat Space Environ Med 63:203-205
Shupak A, Doweck I, Nachtigal D, Spitzer 0, Gordon CR (1993)
Vestibular and audiometric consequences of blast injury to the
ear. Arch Otolaryngol Head Neck Surg 119:1362-1367
Sulaiman ZM, Pilmanis AA, O'Connor RB (1997) Relationship
between age and susceptibility to altitude decompression sickness. Aviat Space Environ Med 68:695-698
Weien RW, Baumgartner N (1990) Altitude decompression sickness: hyperbaric therapy results in 528 cases. Aviat Space
Environ Med 61:833-836
Wicks RE (1989) Alternobaric vertigo: an aeromedical review.
Aviat Space Environ Med 60:67-72

latrogenic vestibular disorders

Iatrogenic vestibular disorders may result as inadvertent effects of the management of dizzy patients
or patients with disorders unrelated to the vestibular
system. The word "iatrogenic" is commonly used
incorrectly in the narrower context of a druginduced condition. This is an error since the word
iatros does not mean a "drug". It means a "physician",
and when combined with the word genesis, the term
is correctly applied to any symptom or condition
created by the physician (Ballantyne 1970). In modern usage, this includes the surgeon. Iatrogenic
vestibular disorders include both the avoidable and
the unavoidable consequences of

drug treatment,
physical therapy, or
surgical procedures.

In the broadest sense, they mayaiso be the result of


medical counsel to the patient or the patient's failure
to comply with therapy.
The following examples of iatrogenic vestibular
disorders do not exhaust a11 the possibilities:

drug treatment
Dizziness and postural imbalance are among the
most frequent adverse effeets of an impressive
list of drugs described elsewhere (Chap. 28). In
addition to dosage-dependent reversible side
effeets, there is also irreversible vestibular loss
eaused by, e.g. gentamicin or quinine.
physical therapy
Chiropractic neck manipulations may eause vertebral artery dissection with subsequent PICA
infarction or thrombosis of the basilar artery.
Physicalliberatory man oeuvres for treatment of
posterior eanal benign paroxysmal positioning
vertigo may lead to transitions to anterior or
horizontal canalolithiasis.
surgical procedures
Inner ear surgery and neurosurgery of the cerebellopontine angle may damage the labyrinth or
the vestibular nerve. Benign paroxysmal posi355

tioning vertigo may be caused by ear or head


surgery. Selective vestibular neurectomy or plugging of the semicircular canals ean eause ehronie
functional vestibular insufficieney.
medical counsel
A simple prescription of bedrest for a nonvestibular disorder can in the susceptible patient
lead to benign paroxysmal positioning vertigo.
avoidable vestibuLar disorders
Early identifieation of a vertebral artery dissection foUowing minimal neck trauma may prevent
subsequent Wallenberg's syndrome or basilar
artery thrombosis if treated immediately with
antieoagulants. Likewise early immunosuppressive therapy for Cogan's syndrome ean prevent
bilateral vestibular failure. Avoidanee of exposure to ehronie high levels of noise prevents not
only noise-induced hearing impairment but also
vestibular defieieney in industrial workers or
military personnel.

Intratympanic gentamicin in
Meniere's disease: desired and
undesired effects
Intratympanie injeetions of gentamicin are increasingly used to treat intraetable vertigo or drop attaeks
due to Meniere's disease (p. 91). Obviously it is possible to damage selectively the dark eells of seeretory
epithelium (and thereby improve endolymphatic
hydrops) before signifieantly affecting vestibular and
cochlear funetions. In fact, treatment is often effective even if it does not impair caloric responses.
However, Magnusson et al. (1991) warned that the
onset of ototoxie effects of gentamicin is delayed.
Oxidative damage of the mitochondria (Hutehin and
Cortopassi 1994) and bloekage of trans duc ti on
ehannels of hair cells (Kroese et a1. 1998) may trigger
hair eell death.

Vertigo

356

These multistage mechanisms are consistent with


the fact that gentamicin ototoxicity is reversible at
an early stage but can become irreversible at a late
stage (Halmagyi et al. 1994). After 1 to 4~ years of
treatment 5 of 18 patients developed oscillopsia and
ataxia, symptoms and signs of (presumably perman ent) chronic vestibular insufficiency (Murofushi et
al. 1997). Responses of the vestibulo-ocular reflex
following unilateral vestibular deafferentation by
gentamicin ablation showed a strong asymmetry
with permanent, reduced gains toward the side of
the lesion (Allison et al. 1997). The proportion of
chronic vestibular insufficiency following this kind
of treatment was not lower than that after selective
vestibular neurectomy or surgicallabyrinthectomy.
Therefore, earlier procedures involving serial
intratympanic injections of gentamicin have been
revised. The current tendency is to treat these
patients with low concentrations of gentamicin and
one injection a week on an outpatient basis, while
carefully monitoring audiovestibular function and
suppression of the distressing attacks (p. 91).

Quinine: reversible and irreversible


side effects
The most common drug used for cerebral malaria is
quinine (Panisko and Keystone 1990; Tracy and
Webster 1996). Serious side effects include cardiac
dysrhythmias, hypoglycaemia, arterial hypotension,
epileptic seizures, agranulocytosis, thrombocytopenia,
anaphylactic reactions, headache, nausea, and most relevant to the present discussion - multisensory loss. Rarely patients with quinine intoxication
may present with reversible or irreversible blindness, hearing loss, and bilateral vestibular failure
(Strupp et al. 1998). The maximum daily dose of
quinine is 1500-1800 mg. Its side effects often limit
dose and duration of its use in therapy (Krogstad
and Herwald 1988; Tracy and Webster 1996; Roden
1996).

Quinine is also used to treat arrhythmia (Roden


1996) and to prevent painful nocturnal muscle
cramps (Jansen et al. 1997). Hearing impairment
(Alvan et al. 1991) and electronystagmographic
abnormalities (Zajtchuk et al. 1984) are due to high

plasma quinine concentrations but have also been


demonstrated in long-term, low-dose quinine ingestion. Two mechanisms are causative: the reversible
blockade of potassium and sodium ion channels
(Roden 1996) and the irreversible deficits due to
vasoconstriction (Smith et al. 1985; Dyson et al. 1985;
Tracy and Webster 1996). In cases of quinine over-

dosage, forced diuresis or plasmapheresis theoretically have only limited efficacy (Bateman et al. 1991;
Dyson et al. 1985; Strupp et al. 1998), since up to
80-90% of these alkaloids are metabolized in the
liver (Roden 1996). In acute intoxications the continuous monitoring by electrocardiogram and laboratory tests is mandatory.

Vertebral artery dissection due to


chiropractic neck manipulation
A survey of 84 cases of vertebral artery dissection
reported minor traumas as the main cause; 70% of
these patients were in their third or fourth decade of
life (Shimizu et al. 1992). The probable traumas preceding the dissection were neck manipulations,
especially chiropractic procedures (52%). The
delayed onset after the neck trauma could take more
than 1 week, but in most cases it occurred within less
than 24 hours. Cervical rotation and extension are
believed to be the causative movement, and the third
segment of the vertebral artery is considered the
most vulnerable. Vertigo may be the first symptom
of vertebral artery dissection after chiropractic neck
manipulations (Vibert et al. 1993). Wallenberg's syndrome (p. 309) is a common complication of vertebral artery dissection.
Violent head movements and sports may cause
vertigo by mechanisms other than that of arterial
dissection. A selective vulnerability of the otoliths
and the organ of Corti has been reported in cases of
audiovestibulopathy induced by high impact aerobics (Weintraub 1994). Since tinnitus, "ear-fullness:'
and sensitivity to barometric pressure change (flying, scuba, swimming) were also present in about
60% of the described 30 individuals, it is more likely
that excessive Valsalva manoeuvres caused perilymph fistula (p. 99) in this subgroup. Perilymph fistula can result from strenuous physical exercise,
heavy lifting or sneezing (Sakikawa et al. 1994),
manipulation of the ear (cleaning cerumen with cotton swabs or sticks (Kubo et al. 1993), or water-jet
irrigation (Wurtele 1981; Anon and Miller 1985).

Surgically induced vestibular


dysfunction
Dizziness and vestibular dysfunction may be caused
by any surgical procedure involving the middle or

latrogenic vestibular disorders


inner ear, the internal auditory canal, or the cerebellar pontine angle. By removing the footplate of the
stapes in patients with otosderosis, stapedectomy
exposes the inner ear; vertigo is almost universal
during the patient's first post-operative days
(Ballantyne and Ajodhia 1984), particularly in the
days of the polyethylene strut (Singleton and Weider
1987). Thirty-three of 54 patients (Ali and Groves
1964) and 28 of 162 otosderotic patients who underwent stapedotomy (Woldag et al. 1995) complained
of vertigo within the first post-operative week. In
about half of these patients the vertigo persisted
after leaving the hospital. High-resolution computed
tomography of the temporal bone allows monitoring
of the operative results. The presence of air bubbles
at the end of the prosthesis is an indirect sign of
perilymphatic fistula (Woldag et al. 1995). When the
surgical technique was changed to indude prosthesis,
fistulas became infrequent. However, perilymphatic
fistula appeared as a complication of tubing or
cochlear implant surgery (Kubo et al. 1993; Huygen
et al. 1994). Bilateral vestibular failure is a major risk
for patients with unilateral vestibular loss who
receive a cochlear implant on the side of preserved
vestibular function.
Therapeutic labyrinthectomy or selective vestibular neurectomy not only cause transient vertigo and
tilts of visual perception of orientation but mayaiso
lead to vertical-oblique diplopia (Safran et al. 1992)
and skew deviation (Riordan-Eva et al. 1997).
Unilateral deafferentation causes a permanent asymmetry of the vestibulo-ocular reflex (Halmagyi 1994;
Allison et al. 1997); approximately 15% of patients
with unilateral vestibular deafferentation develop
oscillopsia and gait ataxia, symptoms of distressing
chronic vestibular insufficiency (Halmagyi 1994). It
could be that vestibular function in the remaining
labyrinth of these patients is not as good as it should
be and that the central compensation for the loss of
vestibular reflexes in response to rapid head movements is normally incomplete (Aw et al. 1994).

latrogenic benign paroxysmal


positioning vertigo
Prolonged bedrest due to unrelated diseases and
post-operative bedrest are well-recognised risk factors for the manifestation of a benign paroxysmal
positioning vertigo (Baloh et al. 1987; Gyo 1988). It is
conceivable that prolonged bedrest with the head
and the inner ear in one position facilitates the formation of a dot, congealed from otoconia debris in
the endolymph, which then becomes symptomatic

357

as canalolithiasis (p. 251). Benign paroxysmal positioning vertigo has also been described as a sequela
of ear or head surgery (Andaz et al. 1993). Here the
most likely explanation is that the operative trauma,
e.g. the removal of an osteoma with hammer and
chisei, caused loosening of otoconia that sub sequently led to canalolithiasis.
Therapeutic physicalliberatory manoeuvres for
typical posterior canal benign paroxysmal positioning vertigo may cause transition of this type into the
anterior or horizontal canal type. Of 85 patients
studied who originally had posterior canal benign
paroxysmal positioning vertigo, 5 had anterior canal
(n = 2) or horizontal canal (n = 3) benign paroxysmal
positioning vertigo after undergoing such manoeuvres (Herdman and Tusa 1997). According to our
experience, these transitions during physical
liberatory manoeuvres for treatment of benign
paroxysmal positioning vertigo (also iatrogenic
transitions from canalolithiasis to cupulolithiasis,
p. 274; Steddin and Brandt 1996) are fortunately only
transient and can be as effectively treated as the
patient's initial condition.
Operative plugging of the posterior semicircular
canal for treatment of benign paroxysmal positioning vertigo is effective, but for more than 99% of the
patients it is unnecessary, since physical therapy is
so successful (p. 265). Possible complications are
surgicallabyrinthitis with reversible ataxia and sensorineural hearing loss or inadvertent plugging of
neighbouring, e.g. horizontal, semicircular canals. In
monkey experiments a direction-changing positional
nystagmus continued for several months after vertical canal plugging (Arai et al. 1989).
Finally, microvascular decompression procedures
for treatment of vestibular paroxysmia ("disabling
positional vertigo", p. 117) can in rare cases damage
the eighth nerve by stretching due to the retraction
of the cerebellum or by heat from electrocoagulation
dose to the nerve. A delayed and progressive hearing
loss can in exceptional cases also result from the formation of reactive scar tissue following a microvascular decompression operation (Fuse and M0ller
1996).

Vestibular loss associated with


chronic noise-induced hearing loss
Industrial workers and military personnel suffering
from noise-induced hearing loss have been shown to
have a symmetrical and centrally compensated
decrease in the vestibular end-organ response, which
is associated with the symmetrical hearing loss

358

(Shupak et al. 1994). Histological and functional


derangements of the vestibular system have also
been reported to occur in the guinea-pig (McCabe
1958; Mangabeira-Albernaz et al. 1959) and the chinchilla (Lim et al. 1982) when the animals were
exposed to high levels of noise. Vestibular malfunction related to noise-induced hearing loss obviously
has some clinical significance beyond the medicolegal implications (Shupak et al. 1994).

References
Ali Y, Groves J (1964) Vestibular disorders after stapedectomy. J
Laryngol Otol 78: 1102-1113
Allison RS, Eizenman M, Tomlinson RD, Nedzelski J, Sharpe JA
(1997) Vestibulo-ocular reflex deficits to rapid head turns following intratympanic gentamicin instillation. J Vestib Res
7:369-380
Alvan G, Karlsson KK, Hellgren U, Villen T (1991) Hearing impairment related to plasma quinine concentration in healthy volunteers. Br J Clin PharmacoI31:409-412
Andaz C, Whittet HB, Ludman H (1993) An unusual cause of
benign paroxysmal positional vertigo. J Laryngol Otol
107:1153-1154
Anon JB, Miller GW (1985) Perilymph fistula. South Med J
78:1454-1457
Arai Y, Henn V, Boehmer A, Suzuki J (1989) How could canalpluggings result in intensive direction changing type of positional nystagmus? Acta Otolaryngol (Stockh) Suppl
468:159-164
Aw ST, Halmagyi GM, Curthoys IS, Todd MI, Yavor RA (1994)
Unilateral vestibular deafferentation causes permanent impairment of the human vertical vestibulo-ocular reflex in the pitch
plane. Exp Brain Res 102: 121-130
Ballantyne JC (1970) Iatrogenic deafness. J Laryngol Otol
84:967-1000
Ballantyne J, Ajodhia J (1984) Iatrogenic dizziness. In: Dix MR,
Hood JD, (eds) Vertigo. Wiley, Chichester, pp 217-247
Baloh RW, Honrubia V, Jacobson K (1987) Benign positional vertigo:
clinical and oculographic features in 240 cases. Neurology
37:371-378
Bateman DN, Blain PG, Woodhouse KW, Rawlins MD, Dyson H,
Heyworth R, Prescott LF, Proudfoot AT (1991)
Pharmacokinetics and clinical toxicity of quinine overdosage:
lack of efficacy of techniques intended to enhance elimination.
Q J Med 54:125-131
Dyson EH, Proudfoot AT, Bateman DN (1985) Quinine amblyopia:
is current management appropriate? J Toxicol Clin Toxicol
23:571-578
Fuse T, M011er MB (1996) Delayed and progressive hearing loss
after microvascular decompression of cranial nerves. Ann Otol
Rhinol Laryngol 105: 158-161
Gyo K (1988) Benign paroxysmal positional vertigo as a complication of postoperative bedrest. Laryngoscope 98:332-333
Halmagyi GM (1994) Vestibular insufficiency following unilateral
vestibular deafferentation. Aust J Otolaryngol 1:510 -5 12
Halmagyi GM, Fattore CM, Curthoys IS, Wade S (1994)
Gentamicin vestibulotoxicity. Otolaryngol Head Neck Surg
111:571-574
Herdman SJ, Tusa RJ (1997) Complications of the canalith repositioning procedure. Arch Otolaryngol Head Neck Surg
122:281-286

Vertigo
Hutchin T, Cortopassi G (1994) Proposed molecular and cellular
mechanism for aminoglycoside ototoxicity. Antimicrob Agents
Chemother 38:2517-2520
Huygen PL, van den Broeck P, Admiraal RJ (1994) Does intracochlear implantation jeopardize vestibular function? Ann Otol
Rhinol LaryngolI03:609-614
Jansen PH, Veenhuizen KC, Wesseling AI, de Boo T, Verbeek AL
(1997) Randomised controlled trial of hydroquinine in muscle
cramps. Lancet 349:528-532
Kroese ABA, Das A, Hudspeth AJ (1989) Blockage of transduction
channels of hair cells in bullfrog's sacculus by aminoglycoside
antibiotics. Hear Res 37:203-218
Krogstad DJ, Herwald BL (1988) Chemoprophylaxis and treatment of malaria. N Engl J Med 319:1538-1540
Kubo T, Kohn M, Naramura H, Jtoh M (1993) Clinical characteristics and hearing recovery in perilymphatic fistulas of different
etiologies. Acta Otolaryngol (Stockh) 113:307-311
Lim DJ, Dunn DE, Johnson DL, Moore TJ (1982) Trauma of the ear
from infrasound.Acta Otolaryngol (Stockh) 94:213-231
McCabe BF, Lawrence M (1958) The effects of intense sound on
the non-auditory labyrinth. Acta Otlaryngol (Stockh)
49:147-157
Magnusson M, Padovan S, Karlberg M, Johansson R (1991)
Delayed onset of ototoxic effects of gentamicin in treatment of
Meniere's disease. Acta Otolaryngol (Stockh) Suppl
481:610-612
Mangabeira-Albernaz PL, Covell WP, Eldredge DH (1959)
Changes in vestibular labyrinth with intense sound.
Laryngoscope 69:1478-1493
Murofushi T, Halmagyi GM, Yavor RA (1997) Intratympanic gentamicin in Meniere's disease: results of therapy. Am J Otol
18:52-57
Panisko DM, Keystone JS (1990) Treatment of malaria - 1990.
Drugs 39:160-189
Riordan-Eva P, Harcourt JP, Faldon M, Brookes GB, Gresty MA
(1997) Skew deviation following vestibular nerve surgery. Ann
NeuroI41:94-99
Roden DM (1996) Antiarrhythmic drugs. In: Hardman JG,
Limbird LE (eds), Goodman & Gilman's, The pharmacological
basis of therapeutics. McGraw-Hili, New York, pp 839-874
Safran AB, Hasler R, Issoua D, Stepanian E, Chiari M, Vibert D,
Roth A (1992) Strabismes induits par des lesions vestibulaires
peripheriques. Klin Monatsbl Augenheilkd 200:418-420
Sakikawa Y, Kobayashi H, Nomura Y (1994) Changes in cerebrospinal fluid pressure in daily life. Ann Otol Rhinol Laryngol
103:959-963
Shimizu J, Nakagawa Y, Fuji Y, Nakase H, Mannen T (1992)
Wallenberg's syndrome due to vertebral artery dissection following minimal neck injury - report of two cases. Rinsho
Shinkeigaku 32:430-435
ShupakA, Bar-EI E, Podoshin L, Spitzer 0, Gordon CR, Ben-David
J (1994) Vestibular findings associated with chronic noise
induced hearing impairment. Acta Otolaryngol (Stockh)
114:579-585
Singleton G, Weider D (1987) Panel discussion: Perilymphatic fistula. Am J OtoI8:355-363
Smith DI, Lawrence M, Hawkins JE jr (1985) Effects of noise and
quinine on the vessels of the stria vascularis: an image analysis
study. Am J OtolaryngoI6:280-289
Steddin S, Brandt Th (1996) Horizontal canal benign paroxysmal
positioning vertigo. (h-BPPV): transition of canalolithiasis to
cupulolithiasis. Ann NeuroI40:918-922
Strupp M, Suckfll M, Schwark Ch, Knig G, Brandt Th (1998)
Akute Chinin-Intoxikation: blind, taub und schwindlig. Mnch
med Wochenschr 140:79-81
Tracy JW, Webster LT (1996) Drugs used in the chemotherapy of
protozoal infections. Malaria. In: Hardman JG, Limbird LE
(eds), Goodman & Gilman's, The pharmacological basis of therapeutics. McGraw-Hili, New York, pp 965-985

latrogenic vestibular disorders


Vibert D, Rohr Le Floch J, Gauthier G (1993) Vertigo as manifestation
of vertebral artery dissection after chiropractic neck manipulations. ORL J Otorhinolaryngol Relat Spec 55:140-142
Weintraub MI (1994) Vestibulopathy induced by high impact
aerobics. A new syndrome: discussion of 30 cases. J Sports Med
Phys Fitness 34:56-63
Woldag K, Meister EF, Kosling S (1995) Diagnosis in persistent
vertigo after stapes surgery. Laryngorhinootologie 74:403-407

359
Wurtele P (1981) Traumatic rupture of the eardrum with round
window fistula. J OtolaryngollO:309-312
Zajtchuk JT, Mihail R, Jewell JS, Dunne MI, Chadwick SG (1984)
Electronystagmographic findings in long-term low-dose
quinine ingestion. A preliminary report. Arch Otolaryngol
110:788-791

SECTION G
Hereditary vestibular disorders
and vertigo in childhood

This section deals with inherited channelopathies


that cause familial episodic ataxia with or without
vertigo (Chap. 25), congenital or early acquired
vestibular dysfunction, and the differential diagnosis
of vertigo in childhood (Chap. 26). Numerous rare
hereditary dis orders can cause peripheral vestibular,
vestibulocochlear (Huygen and Verhagen 1994), or
central vestibular, vestibuloauditory, and oculomotor
dysfunctions (Verhagen and Huygen 1994). Bilateral
vestibular failure (see also Chap. 8) is a typical manifestation, but unilateral vestibular loss or episodic
vertigo may also occur. Although vertigo is less common in the paediatric and adolescent age groups
than in adults, its variety of presentation and under-

lying causes confront the physician with a considerable diagnostic challenge (Chap. 26).

References
Huygen PLM, Verhagen WIM (1994) Peripheral vestibular and
vestibulo-cochlear dysfunction in hereditary dis orders. J Vestib
Res 4:81-104
Verhagen WIM, Huygen PLM (1994) Central vestibular, vestibuloacoustic and oculomotor dysfunction in hereditary disorders. J
Vestib Res 4:105-135

363

Familial periodic ataxia/vertigo


(episodic ataxia)

In 1946 Parker first described a syndrome of periodic


ataxia in 11 patients: in four patients the ataxia was
familial and in the remaining seven, a manifestation
of multiple sclerosis. Ihis was before it had been
recognised that periodic attacks of dysarthria and
ataxia (see Chapters 2 and 15; p. 242) are characteristic paroxysms caused by demyelinated pathways
within pontine and brachium conjunctivum structures in multiple scIerosis (Andermann et a1. 1959;
Harrison and McGill 1969; Osterman and
Westerberg 1975).
Familial periodic ataxia (without MS) is arare,
disabling condition of autosomal dominant inheritance, but it is not a distinct clinical entity.
Synonyms are autosomal dominant episodic ataxia
(EA) or hereditary paroxysmal cerebellar ataxia.
Family members have a similar c1inical syndrome;
however, the syndrome varies considerably from
family to family. Ihis suggests separate disorders
that probably have distinct causes. At least two
groups of disorders have been separated c1inically
(Brunt and van Weerden 1990; Gancher and Nutt
1986; Griggs and Nutt 1995; Brandt and Strupp
1997):
1. episodic ataxia - type 1 (EA-l), which manifests

without vertigo and is associated with "interictal" myokymia, and


2. episodic ataxia - type 2 (EA-2), which often
manifests with vertigo and is associated with
"interictal" nystagmus.
The diagnosis of EA-l and EA-2 is important, since
they can be easily treated and are often mislabeled.
EA-l and EA-2 have been identified to be channelopathies. EA-1 was found to be due to different
heterozygous missense point mutations in a voltagegated (delayed rectifier) potassium channel gene
(KCNA1/Kv1.1) found on chromosome 12p13
(Browne et al. 1994) in seven families, whereas EA-2
is caused by mutations of the cerebral P/Q type calcium
channel a subunit gene CACNLlA4, localised on
chromosome 19p (Vahedi et al. 1995; Browne et al.

1995), which is highly expre sed in the cerebellum.


Here too there were also good reasons to postulate
an ion-channel gene defect, since all autosomal
dominant (acetazolamide-responsive) episodic disorders - such as myotonias or dyskalemic periodic
paraly es (McArdle, 1962; Ptacek et a1. 1994) - have
been caused by mutations in ion-channel genes
(Cannon 1996; Hudson et al. 1995; Griggs and Nutt
1995). Clinically EA-l and EA-2 differ in duration of
the attacks, provoking factors, associated findings,
and progression of the "interictal" signs (Table 25.1).
As effective as acetazolamide is in preventing
attacks, prospective studies still have to prove
whether it can prevent progressive ataxia in EA-2 or
even improve chronic cerebellar deficits. Ihe mode
of action of acetazolamide involves the correction of
an abnormally elevated pH level in the cerebellum
(Bain et al. 1992). There are fascinating similarities
between episodic ataxias and familial hemiplegic
migraine and CADASIL (cerebral autosomaldominant arteriopathy with subcortical infarcts and
leukoencephalopathy), all of which have been
localised on chromosome 19 (Dichgans et a1. 1996).
The three allelic disorders EA-2, familial hemiplegic
migraine, and dominantly inherited late-onset spinocerebellar ataxia type 6 (SCA6) have overlapping
clinical features (Jen et al. 1998).
In summary, familial episodic ataxias are autosomal dominant, inherited disorders caused by
defects of chromosome 12p or 19p; one has been
identified to be a potassium-channel gene defect, the
other a calcium-channel gene defect. Further studies
are necessary to und erstand the triggers for the
attacks (their various duration), the different
involvement of cerebellar structures and the peripheral neuromuscular system, and the progressive cerebellar degeneration in some of the affected families.
The most relevant differential diagnoses for
familial periodic ataxias (EA-l and EA-2) are paroxysmal dysarthria and ataxia in multiple scIerosis
(p. 242), basilar migraine (p. 329), and their variants
of benign paroxysmal vertigo in childhood and
benign recurrent vertigo in adults. A positive family
365

Vertigo

366

history and responsiveness to acetazolamide trongly


upport the diagnosis (Table 25.1).
A number of autosomal or x-linked recessive
metabolic defects can also manifest with episodic
ataxia. These include Hartnup di ease, pyruvate
decarboxylase deficiency, pyruvate dehydrogenase
deficiency, and maple syrup urine disease (Baron et
al. 1956; Blass et al. 1971; Dancis et al. 1967; Farmer
et al. 1973; Lonsdale et al. 1969), all ofwhich should
be treated as entities clearly distinct from autosomal
dominant EA-l and EA-2, which will be discussed here.
Table 25.1. Autosomal dominant episodic ataxias
Disorder

EA-1

EA-2

Chromosome
localisation
Gene lesion

12p13

19p

Potassium channel gene


KCNA l/Kvl.l:
7 missense point
mutations
in 7 families

Calcium channel
gene
CACNLlA4:
2 mutations
disrupting
the reading frames
Hours to days

Duration of attack
(typicalj
Provoking factors

Startle, exercise

On set of attacks

Early childhood

Associated findings

Myokymia; joint
contractu res;
paroxysmal kinesigenic
choreoathetosis

Progressive cerebellar
signs
Treatment

None

Seconds to minutes

Stress, exercise,
fatigue
Adolescence; late
childhood
Nystagmus;
headache;
developmental delay;
cerebellar vermian
atrophy
Frequent

Acetazolamide; phenytoin Acetazolamide

EA-l =episodic ataxia-type 1; EA-2 =episodic ataxia-type 2.


Modified from Griggs and Nutt (1995) and Ophoff et al. (1996).

The clinical syndromes


It is easy to distinguish the two major groups of

autosomal dominant episodic ataxias by their characteristic clinical differences. The key symptoms of
EA-l are short paroxysms of ataxia (mostly without
vertigo) and "interictal" continuous motor unit
activity similar to myokymia and neuromyotonia.
EA-2 shows more heterogeneity among afflicted
families. Key symptoms are Ion ger-lasting paroxysms of ataxia (mostlywith vertigo) and "interictal"
pathological nystagmus.
As indicated in Table 25.1, the differential diagnosis is based mainly on the duration of attacks (seconds to minutes for EA-l ; hours to days for EA-2),

provoking factors (startle, exercise for EA-l; stress,


exercise for EA-2), and associated signs (myokymia
in EA-l; nystagmus and cerebellar signs in EA-2)
(Brunt and van Weerden 1990; Griggs and Nutt
1995).

Episodic ataxia associated with "interictal"


myokymia (type EA-')
Seven families have been described (Van Dyke et al.
1975; Hanson et al. 1977; Gancher and Nutt 1986;
Brunt and van Weerden 1990; Browne et al. 1994,
1995) who have a syndrome combined with central
and peripheral signs classified as EA-l but also
resembling paroxysmal kinesigenic choreoathetosis,
myokymia, and neuromyotonia. Brunt and van
Weerden (1990) have carefully analysed the different
case histories and from their work the following
pattern of attacks emerges (Table 25.2):
"The attacks usually commense between the 5th
and 7th years of life. The onset is sudden and, in
about half of the patients, marked by a click or
shock. Within a few seconds this is followed by a diffuse evolving or spreading sensation of stiffness or
limpness, causing awareness of an attack without the
occurrence of actual movement. The intensity peaks
in a matter of seconds and subsequently declines
gradually over several minutes. In addition to the
characteristic incoordination, most patients notice
trembling of the face, head or limbs and many report
visual blurring and staring of the eyes. A minority
report a feeling of warmth or increased perspiration.
Nausea or rotatory vertigo and involuntary jerking
or writhing movements are absent."
The attacks are frequently elicited by a sudden or
unexpected movement, as in paroxysmal kinesigenic
choreoathetosis. Typical triggers include rising from
achair, running, or performing gymnastics or
sports. During the attacks patients have disturbed
limb coordination and reduced speed of alternating
movements; also their gait is broad-based. Other
typical signs during the attack are postural tremor of
the arms and head, and dysarthria. Myokymia particularly in the face - may be increased.
Myokymia can be observed in some patients
inbetween the attacks, but it is regularly detected by
electromyography as a continuous motor unit activity
with repetitive single or group discharges (duplets
and multiplets) (Andermann et al. 1961; Albers et al.
1981; Brunt and van Weerden 1990). Brunt and van
Weerden (l990) were able to localise the origin of
myokymia in EA-l at the peripheral nerves. Between
the attacks a slight postural tremor and ataxia were
found in a few of the elderly affected family members (Brunt and van Weerden 1990), but progressive
cerebellar atrophy is not typical for EA-l.

367

Familial periodic ataxia/vertigo (episodic ataxia)

Table25.2. Hereditary paroxysmal ataxia and myokymia/"episodic ataxia-type 1"


Van Dyke et al. (1975)

Hanson et al. (1977)

Gancher and Nutt (1986)

Brunt and van Weerden (1990)

Inheritance

AD

AD

AD

AD

Number
reported
examined
affected

3
7
11

3
3
4

3
7
15

22
22
28

2-12

4-8

0-10
3min

0-1
10-15 min

6-15
>20
0-2
2min

2-15
20-50
0-15
10s-10min

Subjective sensations

Sensory warning +
vertigo, dizziness,
blurred vision, diplopia

Sensory warning?
weakness of legs,
eyes drawn backward

Sensory warning?
weightlessness,
sensation offalling

Sensory warning?
limp, stiffness, heavy feeling,
trembling, blurred version

Movements
Coordination

Generalised ataxia, jerking,


nodding of head and arms

Generalised ataxia,
shaking, staggering

Generalised ataxia,
tremor?

Generalised ataxia,
postural tremor head/arms

Stiffening

Generalised myokymia, carpal


spasm

Stiffening and cramp in


body, arms and hands

Twitching in limb muscles, Sometimes stiffening in the


puckering mouth
hands

Provocation

Kinesigenic + caloric stim. +

Kinesigenic + caloric stim. + Kinesigenic? caloric stim. ? Kinesigenic + caloric stim. ?

Additional influences

Hunger, fatigue, excitement,


anxiety

Perimenstrual, illness

Exercise, stress, trauma,


fatigue, excitement

Startle, anxiousness, illness

Prevention byavoiding
stimuli

Yes

Unknown

Unknown

Yes

Ataxia
Age of onset (years)
Attenuation
Frequency/days
Duration

Interval examination
Nystagmus
Ataxia

: few, > 50 years

Myokymia

12 yrs face, hands, moving


fingers, carpal spasm, calf
hypertrophy

Adult: resting tremor in body Adolescence: face, distal


and hands; Children: slight limbs, moving fingers
finger tremor

> 20 years in most: hands,


face; In some children: pending
finger movements

Contractures

None

Neonatal: feet, hands

Foot abnormality:
shortened tendon

None

EMG

CMUA

CMUA, repetitive normal/


polyphasic and grouped
discharges

Neuromyotonia:
repetitive grouped
discharges, rhythmic
singlets

All examinations:
irregular/regular
repeated singlets/multiplets

ataxia: '>
myokymia: ---ataxia: ?,
myokymia: ?

ataxia: '>
myokymia: ---ataxia: ?,
myokymia: ?

ataxia: ---myokymia: '>


ataxia: ---myokymia: ?

ataxia: ---myokymia: ?
ataxia: '>
myokymia:/

normal
type I predominance
axonalloss
peripheral neuropathy

CK elevated
some small fibres
n.d.

normal
neurogenic change
normal

normal
normal (H-E stain)
normal

n.d.

n.d.

n.d.

normal

Treatment
Phenytoin
Acetazolamide
Investigations
Laboratory
Muscle biopsy
Nerve biopsy
Brain microscopy

- = absent; AD = autosomal dominant; CMUA = continuous motor unit activity; n.d. = not done; ---- = no change; '> = decrease; / = increase. From
Brunt and van Weerden (1990).

Episodic ataxia associated with "interictal"


nystagmus (type EA-2)
About 20 families have been described (Parker 1946;
White 1969; Sogg and Hoyt 1962; Farmer and
Mustian 1963; Hill and Sherman 1968; La France et

al. 1977; Donat and Auger 1979; Baloh and Winder


1991; Hawkes 1992; Van-Bogaert et al. 1993) as a
group, but they most probably represent heterogeneous (Baloh et al. 1997) and even genetically distinct
entities (Damji et al. 1996).
The condition commonly manifests in several

368

Vertigo

family members either as recurrent attacks of


unsteadiness of gait and stance or as attacks of vertigo
and nystagmus. Ataxia may involve vestibulocerebellar and spinocerebellar postural imbalance or
neocerebellar dyscoordination of the limbs. Vertigo,
when present, may be rotatory (Farris et al. 1986),
but it is more frequently described as a kind oflightheadedness, to-and-fro dizziness, spatial disorientation, or as positional vertigo. It may involve
downbeat/upbeat nystagmus, skew deviation, gazeevoked nystagmus, positional nystagmus, diplopia,
nausea, and vomiting. Associated symptoms such as
dysarthria, tinnitus, and paraesthesia may occur. The
age of onset of the attacks is inconsistentj it can vary
from early childhood to the sixth decade, even within one family. The condition affects males and
females equaIly. Exertion, fatigue, emotional stress,
or alcohol often precipitate attacks, but kinesigenic

1993

provocation is not typical. Duration of the attacks


may vary from minutes to daysj most often they last
for hours. Attacks may occur daily or be separated by
longer intervals, even years in some cases.
Clinical examination during the attack-free intervals reveals in most families mild nystagmus (especially downbeat nystagmus) and ataxia as weIl as
EEG dysrhythmia. Patients may present with saccadic dysmetria but normal velo city saccades,
impaired ocular pursuit and optokinetic nystagmus,
rebound nystagmus, direction-changing persistent
positional nystagmus, and normal semicircular
canal-ocular reflexes (Baloh et al. 1997j Furman
1997). Ocular motor abnormalities indicate involvement of the vestibulocerebellum and posterior vermis. Figure 25.1 gives an example of an ENG
recording from a patient with EA-2 who initially
(1993) presented with a central vestibular and oculo-

spo ntallcous nystagmus

1997

1993

smooth pU lSll it

1997

a Electronystagmographic (ENG) recordings from a


patient with EA-2 showing spontaneous nystagmus to the right
and horizontal saccadic smooth pursuit to the right.ln 1993, the
now 36-year-old patient initially presented with only central
vestibular and oculomotor disorder showing a spontaneous nystagmus (without episodes of ataxia), horizontal and vertical saccadic smooth pursuit, and dissociated nystagmus to the right. He
developed episodes with ataxia lasting 1-4 days (frequency ca.
1O/year) since 1995. His father also suffers from episodic ataxia.
Clinical examination and ENG recordings did not show any progression of the central vestibular and oculomotor disorder from
1993 to 1997. b Magnetic resonance imaging (Top 1993: T,weighted imaging; Bottoml997: T,-weighted imaging with
0.1 mmol/kg gadolinium) demonstrated atrophy of the cerebellar vermis and cerebral cortex, wh ich also did not progress during
these 4 years. (From Brandt and Strupp 1997.)
Fig.25.1.

369

Familial periodic ataxia/vertigo (episodic ataxia)

motor disorder (without attacks) and 2 years later


developed episodes with ataxia lasting 1-4 days. The
prognosis varies: while some families show spontaneous improvement or occasional remission,
others have periodie ataxia/vertigo followed by slowly progressive (degenerative?) ataxia (Farmer and
Mustian 1963; Farris et al. 1986). Atrophy of the cerebellum or of the cerebellar vermis has been found in
some families (Vighetto et al. 1988). Donat and Auger
(1979) have summarised the clinical spectrum of
familial periodic ataxia now termed EA-2 (Table 25.3).

Differential diagnoses
The most relevant differential diagnoses are basilar
migraine (see Chap. 20, p. 329) and multiple sclerosis
with paroxysmal dysarthria and ataxia (Chap. 2,

p. 29; Chap. 15, p. 242). They can mimic EA-2 in particular, since progressive ocular motor abnormalities
and cerebellar signs may be part of all three disorders. Headache is not a reliable indicator of basilar
migraine, since it accompanies the attack in only
about 70% of cases and is typieally absent in the
juvenile and the adult basilar migraine variants of
benign paroxysmal vertigo in childhood (p. 376)
and benign recurrent vertigo (p. 337). In contrast,
many patients and kindreds with familial periodie
ataxia report considerable headache accompanying
the attack. EA-2 and familial hemiplegie migraine are
possibly allelic disorders (p. 326).
Less relevant differential diagnoses include
vestibular epilepsy (p. 233), periodic intoxication
with drugs or alcohol (p. 395), phobie postural vertigo (p. 469), vestibular paroxysmia (p. 117), Meniere's
disease (p.83), or metabolie diseases such as

Table 25.3. Summary of published reports of fam ilia I periodic ataxia


Parker
(1946)

White
(1969)

No. of patients

23a

Inheritance

"Familial"

Autosomal
dominant

Sogg and Hoyt


(1962)

Farmer and Mustian Hili and Sherman


(1963)
(1968)

La France et al.
(1977)

2a

16a

35a

9a

2a

? Autosomal

Autosomal

Autosomal

Autosomal

? Autosomal

dominant

dominant

dominant

dominant

dominant

Donat and Auger


(1979)

Age ofonset

Adult

Infancy

2 days, 7 months

23-50 years

Infancy

Childhood

9 years; adult

Frequency of
attacks

Daily-monthly

Daily

Daily-yearly

Yearly

Weekly

Daily-monthly

Duration of
attack

Brief

1 h-5 days

!-H h

Fewmin2 months

Month-day

1h-1 week

Precipitating
factors

Fatigue,
emotion

Fatigue, emotion, Fatigue, emotion,


alcohol
looking up

Change in position

Fatigue, emotion,
alcohol

d
e

b
d

b
d

e,c
e,c
c,e

c
d

b
d
b

b
c
b

c,e
d,e
b,c

Symptoms
during attack
Gait ataxia
Limb ataxia
Dysarthria
Vertigo /
dizziness
Nausea
Diplopia
Nystagmus
(oscillopsia)
Tinnitus

e
b
b
d
b

d
b
c
d

b
b
e
b

Examination
between
attacks

Nystagmus,
ataxia

Downbeat
nystagmus

Prognosis

Laboratory
test results
Calorics
b
Urine amino b
acids

Upbeat
nystagmus

Dissociated
nystagmus

Normal

Abate after 30;


b
1-persistent ataxia

Progressive ataxia

Complete recovery b

?Abate

Normal
Normal

Normal
Normal

b
Normal

Normal
Normal

Normal
b

a Each report concerns one family. b Not mentioned; c Prominent feature; d Present; e Absent.
From Donat and Auger (1979).

Ataxia,
nystagmus

b
Normal

Downbeat
nystagmus,
mild ataxia

Vertigo

370

Hartnup disease. If a patient presents with periodic


ataxia with persistent myokymia or mild ataxia and
downbeat nystagmus in the attack-free interval, then
a familial dis order should be suspected and other
family members examined, even if the patient denies
any family history of the syndrome. Diagnosis
of EA-1 and EA-2 is important, since they are treatable conditions which, although rare, are often
mislabelled.

Aetiology and pathomechanism


EA-1 and EA-2 belong -like all autosomal dominant
(acetazolamide-responsive) episodic disorders such
as myotonias or dyskalaemic periodic paralyses
(Ptacek et al. 1994; Lehmann-Horn et al. 1994; Rdel
et al. 1994; Brandt and Strupp 1997; Hurko 1997) - to
the increasing number of the so-called ion-channel
disorders or channelopathies, i.e. diseases caused by
mutations of the sodium, potassium, calcium, or
chloride channel gene. For instance, hyperkalaemic
periodic paralysis, paramyotonia congenita, and the
so-called potassium aggravated myotonias are sodium channelopathies; hypokalaemic periodic paralysis, familial hemiplegic migraine, and EA-2 are
calcium channelopathies; most of the myotonias are
chloride channelopathies; and EA-1 is a potassium
channelopathy (Griggs and Nutt 1995; Ophoff 1996;
Lehmann-Horn et al. 1994; Rdel et al. 1994; Cannon
1996; Hudson et al. 1995; Gutman and Chandy 1993;
Lehmann-Horn and Rdel1996; Hurko 1997). These
disorders show remarkable phenotypic heterogeneity associated with different types and positions of
mutations. Most of these ion-channel gene mutations cause changes in the excitability of the membrane of neurons or muscle cells, and in most
mutations of the sodium and chloride channels the
excitability of the membrane is increased, for
instance, due to an impaired activation of the sodium channel or a decreased chloride conductance of
the cell membrane, which may cause myotonia.

channel gene KCNA1 or Kv1.1 (an intronless gene of


1488 bps which is the homologue of the Shaker gene
of the fruit Hy Drosophila melanogaster) has so far
identified seven different heterozygous missense
point mutations (see Fig. 25.2) in affected individuals
of seven familes, indicating that EA-1 results from
mutations in this gene on chromosome 12p13
(Brown et al. 1994, 1995; Comu et al. 1996).
KCNA1/Kvl.l, a slowly inactivating (delayed rectifier)
potassium channel, is likely to be expressed in the
cerebellum and the peripheral nerve, which may
explain the combination of ataxia and myokymia.
Electrophysiologically, EA-1 is characterised by
episodic failure of excitation of cerebellar neurons leading to the short paroxysms of ataxia - and sustained hyperexcitability of the second motoneurons
- causing "interictal" continuous motor unit activity
similar to myokymia and neuromyotonia. The cellular pathomechanism and structure-function relations of the mutant ion channel in EA-1 have been
studied by patch-clamp recordings of the mutant
potassium channels expressed in Xenopus oocytes
(Adelman et al. 1995): in all of the six analysed mutations the delayed rectifier function was altered (e.g.
faster kinetics, increased "C-type inactivation", shift
of the voltage dependence by 20 mV positive, or
reduction of the potassium current, when coassembled with wild type) in such a way that the affected
nerve cell cannot efficiently repolarise following an
action potential.

Episodic ataxia type-2, a cerebral calcium


channelopathy
Vahedi et al. (1995) and Von Brederlow et al. (1995)
established that EA-2 is localised on chromosome
19p. EA-2 is linked to the same region of chromo-

Episodic ataxia type 1, a potassium


channelopathy
Linkage studies in four families with EA-1 suggested
localisation of the gene dose to several voltage-gated
potassium channel genes (KCNA 1,5, and 6/Kv1.1,
1.5, and 1.6; nomenclature according to the nomendature of mammalian potassium channels by
Gutman and Chandy (1993)) on chromosome 12p13.
Mutation analysis of the voltage-gated potassium

Fig.25.2. Schematic representation of the membrane topology of Kv1.1 subunits. The position of the episodic ataxia type 1
(EA-1) point mutations are indicated by arrows. (From Adelman
et al. 1995.)

Familial periodie ataxia/vertigo (episodie ataxia)

some 19 as is familial hemiplegie migraine and eerebral autosomal-dominant arteriopathy with subcortieal infarcts and leukoencephalopathy (CADASIL)
(Dichgans et al. 1996; Ducros et al. 1996). Reeently
the gene defeet for EA-2 and familial hemiplegie
migraine has been localised on the cerebral (brainspecific) P/Q-type calcium channel a1 subunit gene
CACNL1A4 (Dirlong et al. 1995) (see Fig. 25.3) on
chromosome 19 (Ophoff et al. 1996), which encodes
a protein of 2261 amino acids and is highly
expressed in the cerebellum {Starr et al. 1991).1t is
interesting that a mutation of the calcium channel
gene CACNL1A3, the gene encoding the a1 subunit
of the dihydropyridine receptor, the voltage-gated
skeletal muscle L-type calcium channel, causes
another disease with episodic symptoms: hypokalaemic periodic paralysis (Fontaine et al. 1994;
Jurkat-Rott et al. 1994). Calcium channels are multimeric complexes that are composed of one a1, a2, 6,
and a subunit. The most important subunit is the
a1 subunit, which acts as the voltage sensor and the
ion-conducting pore (for ref. see Guy and Durell
1996). The frame-shift and splice-site mutations in
EA-2, which disrupt the reading frame, predict that it
is unlikely that functional calcium channels are
formed (either the channel assembly is disturbed or
the channel is unstable), so that the density of functioning cerebral PIQ- type channels is decreased
(Ophoff et al. 1996). This may lead to the symptoms
mentioned above.
An ion-channel defect is most likely the cause of
EA-2, because all autosomal-dominant acetazolamide-responsive dis orders to date have been

371

caused by mutations in ion-channel genes (Griggs


and Nutt 1995; Vahedi et al. 1995). There are fascinating similarities between EA-2 and familial hemiplegic migraine. They are both characterised by
attacks, the development of abnormal eye movements, and cerebellar degeneration (Elliot et al.
1996). Nystagmus and ataxia frequently begin prior
to the onset of migraine attacks (Baloh et al. 1994), a
feature that we also observed in our study on basilar
migraine (Dieterich and Brandt 1999). This indicates
that eye movement and cerebellar signs are not
ischaemic sequelae of recurrent migraine but rather
are due to genetically determined cerebellar degeneration (Dieterich and Brandt 1999; Elliot et al. 1996).
Other families have been described with dominantly inherited syndromes different from those
described above. They include familial periodic ataxia
with progressive smooth pursuit defect (Damji et al.
1996), acetazolamide-responsive episodic familial
vertigo (rather than ataxia) with progressive bilateral
vestibular loss (Baloh et al. 1991), and cerebellar
atrophy found only in familial hemiplegic migraine
linked to chromosome 19p13 (Terwindt et al. 1996).
Episodic vertigo is a typical sign in the early stages
of some cases of "idiopathic bilateral vestibular
failure". Familial periodic ataxia of type EA-2 is
obviously a heterogeneous group that requires further elaboration.
There is a precedent in ion-channel diseases for
such remarkable phenotypic heterogeneity being
allelic dis orders associated with different types and
positions of mutations. Patients, for instance, who
have mutations in different portions of the muscle

Domoin

111

IV

Extrocellu'ar

Cytop'asm

EI

R192Q

T666M

.0-

V714A

'*

!'1

deletion C73
spli<8 sil8 mulation

11811L

Fig.25.3. Membrane topology of al subunit of the P/Q-Type Ca 2+-ehannel, CACNL 1A4. The loeation and amino acid substitutions
are indieated for mutations that eause familial hemiplegie migraine (FHM) and episodie ataxia type 2 (EA-2). (From Ophoff et al. 1996.)

372
sodium-channel gene can have widely differing
phenotypes: episodic myotonia without weakness
versus periodic paralysis without myotonia (Griggs
and Nutt 1995; Ptacek et al. 1993a,b, 1994).

Effects of acetazolamide and the


pathomechanism of EA
Acetazolamide's effective prevention of the attacks
via changes in pH (see below) may be one key for
understanding the disease's pathomechanism, since
changes in the intra- and extracellular pH cause
changes of the transmembranous potential and, for
instance, of the potassium conductance: a decrease
in pH reduces, an increase in pH increases potassium
conductance (for ref. see Hille 1993). Nuclear magnetic resonance spectroscopy showed abnormal
intracellular pH levels in the cerebellum of untreated
patients with familial periodic ataxia; these levels
returned to normal with treatment (Bain et al. 1992).
Thus, one can speculate that attacks in familial periodic
ataxia are secondary to abnormally high intracellular cerebellar pH values, and by reducing this
pH level (which reduces potassium conductance of
the cell membrane as mentioned above) acetazolamide prevents attacks. Without treatment, episodes
of cerebellar dysfunction may sporadically occur
with further rises in intracellular pH, which are
sometimes precipitated by exercise or stress (hyperventilation?). The attack is in some way self-limiting
(Bain et al. 1992). Rises in intracellular pH have been
shown to significantly affect both sodium activation
and inactivation as weIl as other biophysical properties of membrane channels, e.g. potassium and calcium
channels (Brodwick and Eaton 1978).
Several reports have discussed the possibility that
enzyme systems such as pyruvate dysmetabolism
are involved in other autosomal or X-linked recessive disorders associated with episodic ataxia (Blass
et al. 1970; Livingstone et al. 1984; Evans et al. 1978;
Sander et al. 1980). Such a connection is supported
by the efficacy of acetazolamide for pyruvate
dysmetabolism (Livingstone et al. 1984; Evans et al.
1978). Explanations of the pathomechanisms of
familial episodic ataxias will have to elucidate both
episodic ataxia and progressive cerebellar degeneration. Selective damage of vestibulocerebellar structures seems to be possible, since the pyruvate
dehydrogenase activity is particularly low at this site
relative to the rate of pyruvate oxidation (Reynolds
and Blass 1976; Zasorin et al. 1983). Vighetto et al.
(1988) tried to explain anterior vermian atrophy in
EA-2 as the result of such a biochemically vulnerable
region. How this is related to blood flow, ischaemia,
and the cerebellar influence on rCBF (regional cerebral blood flow), however, remains uncertain (Elliot
et al. 1996).

Vertigo

Management
First attacks of EA-1 appear in childhood and tend
to abate after early adulthood. Attacks are prevented
first of all by attempting to avoid gymnastics, sports,
and sudden movements (Brunt and van Weerden
1990), which, of course, is not an optimal or even
safe form of management.
Acetazolamide (dosages ranging from 62-750
mg/day) was effective in most patients with ataxia.
About 50% become almost free of attacks (Brunt and
van Weerden 1990), sometimes showing a remarkably long-Iasting improvement following prolonged
treatment with this drug (a similar observation has
been made about the preventative action of beta
blockers in migraine). Acetazolamide has several
effects: it inhibits carbonic anhydrase, thereby
inhibiting the interconversion of CO 2 + H 2 0 to
H ZC0 3; it produces diuresis, initial kaliuresis, metabolic acidosis; it lowers serum bicarbonate levels and
reduces the amount of brain lactate and pyruvate
with subsequent brain acidosis (Zasorin et al. 1983).
Possible adverse effects are nephrocalcinosis, hyperhydrosis, paraesthesia, muscle stiffening with easy
fatigability, and gastrointestinal disturbances. Side
effects are dose related and can be partly reduced by
potassium chloride supplementation (Brunt and van
Weerden 1990).
Of the other carbonic anhydrase-inhibiting drugs,
sulthiame has also been used successfully in EA-1
(Wolf 1980). It caused less side effects and was most
effective in dosages between 50 and 300 mg daily
(Brunt and van Weerden 1990). Although antiepileptic drugs were very effective in paroxysmal
kinesigenic choreoathetosis, they were effective in
only two of the four reported families with EA-1.
Thus, acetazolamide is the drug of first choice. Its
efficacy was discovered by serendipity (Griggs et al.
1978) in a patient with EA-2 misdiagnosed as
periodic paralysis and was later confirmed by Donat
and Auger (1979), Zasorin et al. (1983), and others.
Sulthiame is an alternative anhydrase-inhibiting
drug which effectively prevents the attacks (Wolf
1980). There are only a few reports on the effect of
acetazolamide on "interictal" nystagmus and progressive cerebellar degeneration. In the original
acetazolamide-responsive EA-2 family (Griggs et al.
1978), treatment remained effective for 20 years, and
there has been no progression of ataxia (Griggs and
Nutt 1995). Gancher and Nutt (1986) even described
improvement of"interictal" cerebellar signs.
In view of the proven efficacy of these two potent
drugs, acetazolamide and sulthiame, previous anecdotal reports on other drugs appear to be irrelevant.
Antiepileptic drugs (phenytoin or carbamazepine)

Familial periodic ataxia/vertigo (episodic ataxia)

reduced the frequency of attacks in some patients


(Van Dyke et al. 1975; Hanson et al. 1977) but were
ineffective in others (Gancher and Nutt 1986; Brunt
and van Weerden 1990). Some reduction of symptoms has been attributed to the antivertiginous drug
dimenhydrinate (Farmer and Mustian 1963; Donat
and Auger 1979). However, this kind of symptomatic
treatment cannot be recommended for continuous
administration because of its considerable adverse
sedative effects. Boel and Casaer (1988) successfully
treated a 10-year-old boy with the calcium entry
blocker fiunarizine (10 mg/day).

References
Adelman JP, Bond CT, Pessia M, Maylie J (1995) Episodie ataxia
results from voltage-dependent potassium ehannels with
altered funetions. Neuron 15: 1449-1454
Albers JW, Allen AA, Bastron JA, Daube JR (1981) Limb
myokymia. Muscle and Nerve 4:494-504
Andermann F, Cosgrove JBR, Lioyd-Smith DL (1959) Paroxysmal
dysarthria and ataxia in multiple sclerosis. Neurology 9:
211-215
Andermann F, Cosgrove JBR, Lioyd-Smith DL, Gloor P,
MeNaughton FL (1961) Facial myokymia in multiple sclerosis.
Brain 84:31-44
Bain PG, O'Brien MD, Keevil SF, Porter DA (1992) Familial periodie eerebellar ataxia: a problem of eerebellar intraeellular pH
homeostasis.Ann NeuroI31:147-154
Baloh RW, Winder A (1991) Aeetazolamide-responsive vestibuloeerebellar syndrome: clinieal oculographie features. Neurology
41:429-433
Baloh RW, Jacobson, K, Fife T (1994) Familial vestibulopathy: A
new dominantly inherited syndrome. Neurology 44: 20-25
Baloh RW, Yue Q, Furman JM, Nelson SF (1997) Familial episodie
ataxia: clinieal heterogeneity in four families linked to ehromosome 19p. Ann NeuroI41:8-16
Baron DN, Dent CE, Harris H, Hart EW, Jepson JB (1956)
Hereditary pellagra-like skin rash with temporary eerebellar
ataxia, eonstant renal amino-aeiduria and other bizarre bioehemical features. Laneet 2: 421-428
Blass JP, Avignan J, Uhlendorf BW (1970) Adefeet in pyruvate
deearboxylase in a ehild with an intermittent movement disorder. J Cl in Invest 49: 423-432
Blass JP, Lonsdale D, Uhlendorf BW, Horn E (1971) Intermittent
ataxia with pyruvate-deearboxylase deficieney. Laneet I: 1302
Boel M, Casaer P (1988) Familial periodie ataxia responsive to flunarizine. Neuropediatrics 19:218-220
Brandt Th, Strupp M (1997) Episodie ataxia type 1 and 2 (familial
periodie ataxia/vertigo). Audiol NeurootoI2:373-383
Brodwiek MS, Eaton DC (1978) Sodium ehannel inaetivation in
squid axon is removed by high internal pH or tryosine-specifie
reagents. Scienee 200: 1494-1496
Browne DL, Brunt ER, Griggs RC, Nutt JG, Ganeher ST, Smith EA,
Litt M (1995) Identifieation of two new KCNAI mutations in
episodie ataxia/myokymia families. Hum Mol Genet
4:1617-1672
Browne DL, Ganeher ST, Nutt JG, Brunt ERP, Smith EA, Kramer P,
Litt M (1994) Episodie ataxia/myokymia syndrome is associated
with point mutations in the human potassium ehannel gene,
KCNAl. Nature Genet 8:136-140

373
Brunt ERP, Van Weerden TW (1990) Familial paroxysmal kinesigenie ataxia and eontinuous myokymia. Brain 113: 1361-1382
Cannon SC (1996) Ion-ehannel defeets and aberrant excitability
in myotonia and periodie paralysis. Trends Neurosci 19:3-10
Comu S, Giuliani M, Narayanan V (1996) Episodie ataxia and
myokymia syndrome: a new mutation of potassium ehannel
Gene Kvl. Ann NeuroI40:684-687
Damji KF, Allingham RR, Polloek SC, Small K, Lewis KE, Stajieh
JM, Yamaoka LH, Vanee JM, Periak Vanee MA (1996) Periodie
vestibulo-eerebellar ataxia, an autosom al dominant ataxia with
defeetive smooth pursuit is genetieally distinct from other
autosomal dominant ataxias. Areh Neurol 53:338-344
Dancis J, Hutzier J, Rokkones T (1967) Intermittent branehed
ehain ketonuria: variant of maple-syrup-urine disease. N Engl J
Med 276: 84-89
Diehgans M, Mayer M, Mller-Myhsok B, Straube A, Gasser T
(1996) Identifieation of a key recombinant narrows the
CADASIL gene region to 8eM and argues against allelism of
CADASIL and familial hemiplegie migraine. Genomies
32:151-154
Dieterieh M, Brandt Th (1999) Episodie vertigo related to
migraine (90 eases): vestibular migraine? J Neurol (submitted)
Dirlong S, Lory P, Williams ME, Ellis SB, Harpold MM, Taviaux S
(1995) Chromosomal loealilzation of the human genes for
Alpha lA, IB, and lE voltage-dependent Ca2+ ehannel subunits.
Genomies 30:605-609
Donat JR, Auger R (1979) Familial periodie ataxia. Areh Neurol36:
568-569
Dueros A, Nagy T, Alamowiteh S, Nibbio A, Joutel A, Vahedi K,
Chabriat H, Iba Zizen MT, Julien J, Davous P, et al (1996)
Cerebral autosomal dominant arteriopathy with subeortieal
infarets and leukoeneephalopathy, genetie homogeneity, and
mapping of the loeus within a 2-eM interval. Am J Hum Genet
58:171-181
Elliot MA, Peroutkas J, Welch S, May EF (1996) Familial hemiplegie migraine, nystagmus, and eerebellar atrophy. Ann Neurol
39:100-106
Evans OB, Kilroy AW, Feniehel GM (1978) Aeetazolamide in the
treatment of pyruvate dysmetabolism syndromes. Areh Neurol
35:302-305
Farmer TW, Mustian VM (1963) Vestibuloeerebellar ataxia. Areh
Neurol 8: 471-480
Farmer TW, Veath L, Miller AL (1973) Pyruvate deearboxylase
aetivity in familial intermittent eerebellar ataxia. Trans Am
Neurol Assoe 98: 260-262
Farris BK, Smith JL, Ayyar R (1986) Neuro-ophthalmologie findings in vestibuloeerebellar ataxia. Areh Neurol43: 1050-1053
Fontaine B, Vale Santos J, Jurkat-Rott K, Reboul J, Plassart E, Rime
CS, Elbaz A, Heine R, Guimaraes J, Weissenbaeh J (1994)
Mapping of the hypokalaemie periodie paralysis (HypoPP)
loeus to ehromosome lq31-32 in three European families.
Nature Genet 6:267-272
Furman JM (1997) Otolith -oeular response in familial episodie
ataxia linked to ehromosome 19p. Ann NeuroI42:189-193
Ganeher ST, Nutt JG (1986) Autosomal dominant episodie ataxia:
a heterogeneous syndrome. Mov Disord 1:239-253
Griggs RC, Moxley RT, La Franee RA, MeQuilien J (1978)
Hereditary paroxysmal ataxia: response to aeetazolamide.
Neurology 28: 1259-1264
Griggs RC, Nutt JG (1995) Episodie ataxias as ehannelopathies.
Ann NeuroI37:285-286
Gutman GA, Chandy KG (1993) Nomenclature of mammalian
voltage-dependent potassium ehannel genes. Neuroseienee
5:101-106
Guy HR, DureIl SR (1996) Three-dimensional models of ion ehannel pro teins. Plenum Press, New York
Hanson PA, Martinez LB, Cassidy R (1977) Contraetures, continuous muscle diseharges, and titubation. Ann Neuroll:120-124
Harrison M, MeGili JI (1969) Transient neurological disturbanees

374
in disseminated sclerosis: a case report. J Neurol Neurosurg
Psychiatry 32: 230-232
Hawkes CH (1992) Familial paroxysmal ataxia: report of a family.
J Neurol Neurosurg Psychiatry 55:212-213
Hili W, Sherman H (1968) Acute intermittent cerebellar ataxia.
Arch NeuroI18:350-357
Hille B (1993) Ion channels in exeitable membran es. Sinauer,
Sunderland, Mass
Hudson AI, Ebers GC, Bulman DE (1995) The skeletal muscle
sodium and chloride channel diseases. Brain 118:547-563
Hurko 0 (1997) Recent advances in heritable ataxias. Ann Neurol
41:4-5
Jen JC, Yue Q, Karrim I, Nelson SF, Baloh RW (1998) SCA6 with
positional vertigo and acetazolamide-responsive episodic ataxia. J
Neurol Neurosurg Psychiatry 65:565-568
Jurkat-Rott K, Lehmann-Horn F, Elbaz A, Heine R, Gregg RG,
Hogan K, Powers PA, Lapie P, Vale Santos JE, Weissenbach J
(1994) A calcium channel mutation causing hypokalemic periodic paralysis. Hum Mol Genet 3:1415-1419
La France R, Griggs R, Moxley R (1977) Hereditary proxysmal
ataxia responsive to acetazolamide. Neurology 27:370
Lehmann-Horn F, Engel AG, Ricker K, Rdel R (1994) The periodic
paralyses and paramyotomia congentia. In: Engel AG, FranziniArmstrong C (eds) Myology. McGraw-HilI, New York, pp
1303-1334
Lehmann-Horn F, Rdel R (1996) Molecular pathophysiology of
voltage-gated ion channels. Rev Physiol Biochem Pharamcol
128:198-267
Livingstone IR, Gardner-Medwin D, Pennington RJT (1984)
Familial intermittent ataxia with possible X-linked recessive
inheritance. J Neurol Sei 64:89-97
Lonsdale D, Faulkner WR, Price JW, Smeby RR (1969) Intermittent
cerebellar ataxia assoeiated with hyperpyruvic acidemia,
hyperalaninemia, and hyperalalinuria. Pediatrics 43: 1025-1034
McArdle B (1962) Adynamia episodica hereditaria and its treatment. Brain 85: 121-148
Ophoff RA, Terwindt GM, Vergouwe MN, Vaneijk R, Oefner PI,
Hoffman SMG, Lamerdin JE, Mohrenweiser HW, Bulman DE,
Ferrari M, Haan I, Lindhout D, van Ommen GJB, Hofker MH,
Ferrari MD, Frants RR (1996) Familial hemiplegie migraine and
episodic ataxia type 2 are caused by mutations in the Ca2+
channel gene CACNLlA4. Ce1l87:543-552
Osterman PO, Westerberg CE (1975) Paroxysmal attacks in multiple
sclerosis. Brain 98:189-202
Parker HL (1946) Periodic ataxia. Coll Papers Mayo Clin 38:
642-645
Ptacek LJ, Johnson KI, Griggs RC (1993a) Mechanisms of disease:
genetics and physiology of the myotonie muscle disorders. N
Engl J Med 328:482-489
Ptacek LI, Gouw L, Kwiecinski H, McManis P, Mendall JR, Barohn

Vertigo
RJ, George AL Jr, Barchi RL, Robertson M, Leppert MF (1993b)
Sodium channel mutations in paramyotonia congenita and
hyperkalemic periodic paralysis. Ann NeuroI33:300-307
Ptacek LI, Tawil R, Griggs RC, Meola G, McManis P, Barohn RJ,
Mendell JR, Harris C, Spitzer R, Santiago F, Leppert MF (1994)
Sodium channel mutations in acetazolamide-responsive
myotonia congenita, paramyotonia congenita, and hyperkalemic periodic paralysis. Neurology 44:1500-1503
Reynolds SF, Blass JP (1976) A possible mechanism for selective
cerebellar damage in partial pyruvate defieiency. Neurology
26:625-628
Rdel R, Lehmann-Horn F, Ricker K (1994) The nondystrophic
myotonias. In: Engel AG, Franzini-Armstrong C (eds) Myology.
McGraw-Hill,NewYork,pp 1291-1302
Sander JE, Malamud N, Cowan MJ, Packman S, Amman AJ, Wara
DW (1980) Intermittent ataxia and immunodeficiency with
multiple carboxylase deficiencies: a biotin-responsive dis order.
Ann NeuroI8:544-547
Sogg RL, Hoyt WF (1962) Intermittent vertical nystagmus in a
father and son.Arch OphthalmoI68:515-517
Starr TVB, Prysay W, Snutch T (1991) Primary structure of a calcium
channel that is highly expressed in the rat cerebellum. Proc Natl
Acad Sei USA 88:5621-5625
Terwindt GM, Ophoff RA, Haan J, Frants RR, Ferrari MD (1996)
Familial hemiplegie migraine: a clinical comparison of families
linked and unlinked to chromosome 19. Cephalalgia
16:153-155
Vahedi K, Joutel A, Van Bogaert P, Ducros A, Maciazeck J, Bach JF,
Bousser MG, Tournier-Lasserve E (1995) A gene for hereditary
paroxsymal cerebellar ataxia maps to chromosome 19p. Ann
NeuroI37:289-293
Van Dyke DH, Griggs RC, Murphy MJ (1975) Hereditary
myokymia and periodic ataxia. J Neurol Sei 25: 109-118
Van-Bogaert P, Van-Nechel C, Goldman S, Szliwowski HB (1993)
Acetazolamide-responsive hereditary paroxysmal ataxia: report
of a new family. Acta Neurol Belg 93:268-275
Vi ghetto A, Froment JC, Trillet M,Aimard G (1988) Magnetic resonance imaging in familial paroxysmal ataxia. Arch Neurol
45:547-549
Von Brederlow B, Hahn AF, Koopman WI, Ebers GC, Bulman D
(1995) Mapping the gene for acetazolamide responsive hereditary paroxysmal cerebellar ataxia to chromosome 19p. Human
Mol Genet 2:279-284
White JC (1969) Familial periodic nystagmus, vertigo, and ataxia.
Arch Neurol20: 276-280
Wolf P (1980) Familire episodische Ataxie. Nervenarzt
51:355-358
Zasorin NL, Baloh RW, Myers LB (1983) Acetazolamide-responsive episodic ataxia syndrome. Neurology 33:1212-1214

Vertigo in childhood

Vertigo occurs less frequently in childhood than in


adulthood. If present, however, it can be caused by a
variety of peripheral and central vestibular syndromes that pose a diagnostic challenge (Beddoe
1977; Eviatar and Eviatar 1977; Blayney and Colman
1984; Brandt and Bchele 1984; Balkany and Finkel
1986; Britton and Block 1988; Eviatar 1994; Bower
and Cotton 1995). Inasmuch as most vertigo syndromes in childhood also manifest in adults, they are
described in more detail elsewhere in the relevant
chapter. While considerable effort is expended to
prevent hearing loss in children, concurrent or subsequent vestibular disorders are often ignored
because vertiginous crises in childhood are usually
attributed to problems of behaviour or lack of coordination (Tusa et a1. 1994). As children often lack the
communication skills to give an accurate account of
their speIls, the distinction between vertigo, dizziness and imbalance becomes less obvious (Bower
and Cotton 1995).
Views on the frequency and the most common
causes of vertigo in childhood differ according to
wh ether they originate from children's, otolaryngology,
or neurological departments and whether outpatients or inpatients are evaluated. Harrison (1962)
found a predominance of peripheral causes induding Meniere's disease; in a group of 50 patients who
presented to a neurologicai dinie, Eviatar and
Eviatar (1977) found evidence of central vertigo in
42 and diagnosed vertiginous seizures in half of the
children; Fried (I980) found head trauma to be the
most common cause of vertigo in children, whereas
Bower and Cotton (1995) in their study of outpatients in an otolaryngology dinic found that the
majority of children had vertigo caused by otitis
media, benign paroxysmal vertigo of childhood,
migraine, or vestibular neuritis. Our experience in a
neurological dizziness unit shows that the most
important episodic vertigo syndromes in childhood
and adolescence are benign paroxysmal vertigo
of childhood, a basilar migraine variant (p. 335),
epileptic aura or vestibular epilepsy (p. 233),
375

perilymph fistula (p. 99), "secondary Meniere's dis ease" (p. 83) and familial episodie ataxia (p. 365).
Sustained rotational vertigo due to an acute unilateral vestibular loss may be present with labyrinthitis, vestibular neuritis (p. 67), or foJlowing head
trauma (p. 347). In the latter, post-concussional
otolith vertigo (p. 349) and typical benign paroxysmal
positioning vertigo (p. 251) mayaiso occur.
Oscillopsia with head motion and gait imbalance
in darkness are typical symptoms of bilateral
vestibular failure (p. 127), since it occurs, for example,
with bilateral acoustic neurinoma or as a sequela of
bacterial meningitis, inner-ear autoimmune disease,
or ototoxic drugs. Various rare genetic and embryopathie labyrinth malformations (p. 378) and multiple hereditary diseases can cause congenital as weil
as early acquired unilateral or bilateralloss of auditory and/or vestibular function.
The relatively common infratentorial tumours in
children (p. 244), which typically manifest with fluctuating progressive vertigo, ataxia, and ocular motor
abnormalities, constitute a significant differential
diagnosis for dizziness, vertigo, and disequilibrium
in infancy. Examples of other rare conditions that
cause central vertigo syndromes are congenital or
toxie upbeat/downbeat nystagmus and epidemie
vertigo secondary to viral infections.
To determine the cause of vertigo, the quality and
duration of dizziness or vertigo, associated symptoms, precipitating factors, and events surrounding
the attacks are often more helpful than apparative
diagnostics. The following three key signs and
symptoms should help to make a correct diagnosis:
1. Vertigo attacks
episodic vertigo without "typical" interictal
abnormalities: benign paroxysmal vertigo of
childhood, epileptic seizures, orthostatic
hypotension, psychogenic vertigo
episodic vertigo with sensorineural hearing
loss: perilymph fistula, Meniere's disease,
vestibular paroxysmia

Vertigo

376

episodic vertigo !Vitll "interietal" oeLllomotor


abnormalities: basilar migraine, familial episodic
ataxia
episodic vertigo followed by head motion oscillopsia and gait imbalance in darklless: development of bilateral vestibular failure, familial
vestibulopathy
2. Sustained vertigo
sustained vertigo without hearing loss: vestibular neuritis
sustained vertigo with hearil1g 1055: labyrinthitis,
otitis media, autoimmune disease
post-traumatic sllstained vertigo: temporal
bane fracture, labyrinthine concussion
3. Oscillopsia and postural imbalance
delayed stallce and gait development with or
without hearing loss: congenital bilateral
vestibular failure
head motion oscillopsia and gait im balance in
darkness: congenital or early acquired vestibular fa Hure, perilymph fistula, traumatic otolith
vertigo
sLowly progressive head motion oscillopsia, gait
imbalallce, and sensorinellral hearing 1055:
multiple rare inherited disorders causing progressive audiovestibular loss
progressive ataxia, vertigo, and ocular motor
disorders: infratentorial tumours which affect
vestibulocerebellar or pontomedullary brainstern structures

Benign paroxysmal vertigo of


childhood and basilar migraine
If episodic vertigo occurs in childhood, benign
paroxysmal vertigo of childhood (Basser 1964)
should be considered. It is not a rare disorder, but on
the contrary, probably the most common cause of
episodic vertigo in childhood, occurring with a
prevalence of 2.6% in a population-based study
(Abu-Arafeh and RusseI1995). Relevant differential
diagnoses include epileptic seizures, orthostatic
hypotension, and psychogenic vertigo. Benign
paroxysmal vertigo of childhood is characterised by
sudden brief attacks of incapacitating vertigo associated with nystagmus, onset between ages 1 and 4
years, and spontaneous remission within a few years
(for clinical syndromes and management see Chap.
20, p. 335). Its causative mechanism is presumed to
be equivalent to that of a "migraine", but mostly
without headache. Transitions are possible from this
apparently distinct entity to "common migraine" and

basilar migraine. A long-term follow-up of seven


cases of benign paroxysmal vertigo of childhood
revealed that five resolved spontaneously and six
developed migraine later (Lanzi et a1. 1994).
Benign paroxysmal torticollis in infancy is another
rare migraine variant, which occurs during the first
months of life and resolves spontaneously within 1
to 5 years (Bratt and Menelaus 1992; Cohen et a1.
1993). Basilar artery migraine (Bickerstaff 1961)
typically has visual, vestibular, and other brainstem/cerebellar/temporallobe symptoms. Contrary
to its original description as a dis order of the adolescent girl, it can manifest throughout li fe and in both
sexes (p. 329). The diagnosis is mainly based on a
positive family history and a sequence of events
including recovery and associated headache.

Motion sickness
Children younger than 2 years old are highly resistant to motion sickness (p. 490), but later, until about
the age of 12, they reveal a greater susceptibility than
adults. This has been attributed to their preference
for riding in vehicles in a supine position (Chinn
and Smith 1955), but it can be more convincingly
explained by the finding that dynamic spatial orientat ion in young children does not suffer from the
visual-vestibular mismatches responsible for provoking motion sickness (Brandt et a1. 1976).
Susceptibility to motion sickness also seems to be
higher among sufferers of benign paroxysmal vertigo
of childhood (Abu-Arefeh and Russel 1995), in
children with other forms of migraine (Barabas et a1.
1983), and has also been reported in adult
migraineurs (p. 334).

Vestibular neuritis
In children vestibular neuritis (for clinical syndromes and management, see Chap. 4, p. 67) is a rare
disease, which seems to affect boys more frequently
than girls (Tahara et a1. 1993; Shirabe 1988). As distinct from vestibular neuritis in adults, there is a
high er frequency of preceding upper respiratory
infections. The course is characterised by a more
rapid recovery and better prognosis as to the
restoration rate of labyrinthine function (Shirabe
1988; Tahara et a1. 1993).

Vertigo in childhood

Meniere's disease
Meniere's disease (for clinical syndromes and management, see Chap. 5), with the classic triad of
attacks of vertigo, tinnitus, and ftuctuating hearing
loss (p. 83), is rare in infancy and childhood. Only
1% of cases of Meniere's disease occur in children
under 9 years of age (Harrison 1962). It should be
suspected in infancy whenever vomiting without
diarrhea is associated with some hearing loss (Sade
and Yaniv 1984). Hausler et a1. (1987) described 9 of
14 children aged 14 years or younger as having
"idiopathic Meniere's disease". In 5 children
"secondary Meniere's syndrome" was suspected,
based on histories of an initial hearing loss following
mumps, Haemophilus injluenzae meningitis, temporal bone fracture, or congenital or embryopathic
complications in the affected ear 5 to 11 years prior
to the manifestation of Meniere's syndrome (Hausler
et a1. 1987). Thus, "delayed endolymphatic hydrops"
(p. 88), an impairment of endolymph resorption secondary to labyrinthitis or trauma, and congenital
malformations of the inner ear such as Mondini dysplasia (p. 147), are possible causes of Meniere's disease in infancy. The most important differential
diagnosis is that of a perilymph fistula (p. 99).

Perilymph fistulas
Perilymph fistulas (for clinical syndromes and management, see Chap. 6) are difficult to diagnose
(p. 101). These patients frequently present with combinations of episodic vertigo and hearing loss.
Congenital (Crook 1967; Rice and Waggoner 1967;
Weider and Musiek 1984; Weber et a1. 1993) and
acquired fistulas (Petroff et a1. 1986) should also be
considered in children who have unexplained, sudden, or progressive sensorineural hearing loss, or
complaints of vertigo/hearing loss following trauma
to the ear or head or barotrauma (p. 351) Eighty-six
percent of ears found to have a perilymphatic fistula
had a deformity of the middle ear, inner ear, or both
(Weber et a1. 1993). A malformed stapes was the
most common abnormality, followed by a deformed
round window or a deformed incus. CT or MRI
reveals that these malformations often coexist with
inner-ear malformation (Weber et a1. 1993). The
enlarged vestibular aqueduct syndrome has been
described as an association of an enlarged vestibular
aqueduct, sensorineural hearing loss, round-window
abnormality, and apredisposition to perilymph fistula (Schessel and Nedzelski 1992; Belenky et a1.
1993). This syndrome has a familial incidence, and
the pedigrees of the familial cases show a pattern

377

that seems most consistent with autosomal recessive


inheritance (Tong et a1. 1997). The cochlear aqueduct, more often patent in children than in adults,
may transfer intracranial pressure changes (resulting from violent body exercise) to the labyrinth, and
produce a fistula by explosive damage (p.353).
Surgical tympanoscopy is required to establish the
diagnosis, but the treatment is preferably conservative rather than surgical (p. 105).

Unilateral or bilateralloss of
vestibular function
Causes in children can include congenital malformation of the labyrinth or embryopathic factors, such
as rubella infection during the first trimester of
pregnancy (Barr and Lundstrom 1961), subclinical
cytomegalovirus infection (Pappas 1983), sickle-cell
disease (Savunder et a1. 1996), mitochondriopathy
(Lenard et a1. 1992), or toxins (e.g. thalidomide). The
absence of the bony semicircular canals in the presence of a bony cochlea is a characteristic finding in
the CHARGE (coloboma, heart disease, atresia of
choane, retarded growth and development and/or
central nervous system anomalies, genital hypoplasia, and ear anomalies) association (Murofushi et al.
1997) and was reported in Calman's syndrome (Hill
et a1. 1992). Acute unilateral vestibular loss with sustained rotational vertigo, nystagmus, nausea, and
horizontal semicircular canal paresis may be secondary to viral, bacterial, tuberculous, or syphilitic
labyrinthitis, vestibular neuritis (p. 67), Ramsay
Hunt syndrome, cholesteatoma, trauma, or Cogan's
syndrome, an autoimmune disease that mostly
affects young adults but has also been described in
children (Vollertsen et al. 1986).
Children with congenital or early acquired profound hearing loss often have associated vestibular
loss (Arnvig 1955; Horak et a1. 1988). Their motor
development can be delayed with respect to first
stance and gait in comparison to that of healthy
children (Butterfieid 1986). In later childhood postural performance appears normal (Brunt and
Broadhead 1982; Butterfieid 1986), and conventional
posturography could not distinguish any difference
in postural sway when compared with that of healthy
controls (Enbom et a1. 1991). Body sway increased,
however, when standing on foam rubber, which
reduces the reliability of somatosensory input that
substitutes for the lack of vestibular postural contro1. Accordingly, postural stability is aggravated in
Usher's syndrome due to the associated visual
impairment caused by retinitis pigmentosa (Pyykk
et a1. 1991).

Vertigo

378

The key symptoms of bilateral vestibular failure


are oscillopsia with head motion and unsteadiness
in darkness. Distressing oscillopsia is caused by the
retinal slip of the visual scene consequent to the
defective vestibular ocular reflex. Unsteadiness
occurs if vision or somatosensors cannot compensate (substitute) for the lack of vestibular
information (Chap. 8, p. 127). Gradual bilateral
vestibular failure as it occurs in hereditary disorders
or in aminoglycoside ototoxicity - as long as it is
symmetrical - often remains undetected. To see a
child develop vestibular failure is rare, because
assessment is made only if symptoms necessitate
investigation (Longridge 1989). In idiopathic bilateral vestibular failure (p. 132) and familial vestibulopathy (Baloh et al. 1994), episodic vertigo may
occur in the initial stage of the disease. Acquired
bilateral vestibular failure can be a complication of
meningitis, labyrinthitis, or cerebellopontine angle
tumours (Simmons 1973; Chambers et al. 1985). In a
follow-up study of 94 consecutive survivors of pneumococcal meningitis 4-16 years after dis charge, 17

had hearing loss and l3 vestibular areflexia


(Rasmussen et al. 1991). Audiovestibular sequelae
may even progress during follow-up (Aust 1994).
Table 26.1 summarises the information given in
this chapter about vertigo or loss of vestibular function in children.

Hereditary disorders causing


peripheral vestibular failure
Huygen and Verhagen (1994) have carefully
reviewed the various hereditary dis orders with
peripheral vestibular and vestibulocochlear involvement. These include familial congenital vestibular
areflexia (Verhagen et al. 1987) and dominantly
inherited familial vestibulopathy with normal hearing (Baloh et al. 1994), which present with episodic
vertigo followed by gait imbalance and oscillopsia.
The latter condition possibly responds to aceta-

Table 26.1. Vertigo or 1055 of vestibular funetion in ehildren


Labyrinth/nerve

----

Central vestibular
----

-----------------

Familial/congenital

Labyrinthine malformation (e.g. Mondini dysplasia)

Familial episodic ataxia (EA-2)

Perilymph fistulas embryopathie malformation (rubella,


eytomegalovirus, toxie, e.g. thalidomide)

Downbeat nystagmus
Upbeat nystagmus

Various hereditary disorders with eongenital or


early aequired vestibular 1055
Familial vestibular areflexia
Syphilitie labyrinthitis (endolymphatic hydrops)

Positive migraine family history

Benign paroxysmal vertigo of childhood


Basilar migraine
Benign paroxysmal torticollis of infaney
Acquired
Labyrinthitis with vestibular loss (viral, baeterial,
tubereulous)

Infratentorial tumours (medulloblastoma,


astroeytoma, epidermoid eysts, meningeoma)

Perilymph fistulas
Trauma (temporal bone fraeture, benign paroxysmal
positioning vertigo)

Epileptie aura, vestibular epilepsy


Trauma (brainstem or vestibulo-eerebellar
eoncussion)

Meniere's disease (delayed endolymphatie hydrops)

Eneephalitis

Vestibular neuritis

Toxie (e.g. upbeat/downbeat nystagmus)

Herpes zoster otieus


Cholesteatoma
Ototoxie drugs
Cogan's syndrome
Other inner-ear autoimmune disorders

379

Vertigo in childhood

zolamide and is closely related to migraine.


Congenital or early acquired hearing loss is often
associated with vestibular failure or loss as are
Waardenburg's syndrome and the branchio-otorenal dysplasia syndrome (Huygen and Ver hagen
1994). Bilateral vestibular schwannomas are present
in about 85% of neurofibromatosis type-2 (Gardner
and Frazier 1930). Usher's syndrome, an autosomal
recessive disorder with congenital and usually progressive sensorineural hearing loss, is also characterised by slowly progressive visual loss due to
retinitis pigmentosa. Generally type-1 of Usher's
syndrome also has loss of vestibular function
(Kumar et al. 1984; Trop et al. 1995), and postural
stability is impaired because of the combined lack of
visual and vestibular information (Pyykk et al.
1991). Vestibular hyporeflexia or areflexia is found in
most cases of Friedreich's ataxia (Baloh et al. 1975;
EIl et al. 1984; Cassandro et al. 1986). Furthermore,
vestibular dysfunction has been described for
Anderson-Fabry disease and Alport's syndrome
(Huygen and Verhagen 1994), Stickler's syndrome hereditary arthro-ophthalmopathy - (Vanniasegaram
and Bellman 1994), and bilateral Mondini dysplasia
(Fig. 9.4) with normal hearing (Komune et al. 1993).

Central vestibular syndromes


Cerebellar and brainstem tumours (medulloblastoma, astrocytoma, meningeoma, epidermoid cysts)
are important causes of central vestibular vertigo
(p. 167). They are even more common in children
than in adults. The von Hippel-Lindau disease is an
autosomal-dominant condition with variable penetrance, characterised by midline cerebellar haemangioblastoma and retinal angiomas (Horton et al.
1976). All these tumours affect intra-axial structures
of the vestibulocerebellum and the vestibular nuclei
more often than peripheral nerves. Vertigo is typicaUy progressive (fluctuating in severity but not
episodic) and is associated with ataxia and ocular
motor deficits. All forms of central vestibular syndromes such as positional vertigo, upbeat/downbeat
nystagmus, or ocular tilt reaction may occur. Careful
patient history together with CT and MRI scans will
help to easily distinguish infratentorial tumours
from familial episodic ataxia (p. 365) or basilar
migraine (p. 329), conditions that also present with
ocular motor abnormalities in the vertigo-free
interval.
Sudden prolonged but benign vertigo, not
unequivocally associated with peripheral caloric
hyporesponsiveness, has also been described as

epidemie central vertigo. It may occur after various


viral infections (Mangabeira-Albernaz and
Ganancam 1988).
Downbeat nystagmus (p. 199) is a rare condition
in children. It may be hereditary (e.g. ataxia teleangiectasis), congenital, and persistent (Bixenman
1983). It can also be caused by Arnold-Chiari malformation, infratentorial tumours, or occur in infancy
with spontaneous remission (Weissman et al. 1988).
Congenital upbeat nystagmus is found only rarely
(Shibasaki et al. 1978; Hoyt and Gelbart 1984), but it
is not clear whether this vertical nystagmus is identical to the acquired upbeat nystagmus/vertigo syndrome described elsewhere (p. 205). Upbeat
nystagmus can be part of multiple ocular motor
signs in autosomal dominant early onset nonprogressive cerebellar syndrome (Kattah et al. 1983;
Furman et al. 1986). Periodic alternating nystagmus
is a vestibulocerebellar disorder (Furman et al.
1990), which is congenital in about 25% of cases
(Oosterveld and Rademakers 1979). Familial
vestibulocerebellar dysfunction with episodic
"motion sickness" and "interictal" oculomotor
abnormalities as described by Theunissen et al.
(1989) was considered a new syndrome; however, it
could be a variant of familial episodic ataxia (p. 365).
Various types of cerebellar degeneration can cause
central and/or peripheral vestibular dysfunction
(see review, Verhagen and Huygen 1994), but often
only later in the course of the disease. Central
vestibulocerebellar signs are also part of the multisystem involvement in X-linked ataxia carriers
(Verhagen et al. 1996).

References
Abu-Arafeh I, Russel G (1995) Paroxysmal vertigo as a migraine
equivalent in children: a population-based study. Cephalalgia
15:22-25

Arnvig J (1955) Vestibular function in deafness and severe hardness of hearing. Acta Otolaryngol (Stockh) 4:283-288
Aust G (1991) Gleichgewichtsstrungen und ihre Diagnostik im
Kindesalter. Laryngo-Rhino-Otol 70:532-537
Aust G (1994) Early and late damage to the auditory and vestibular area after meningitis in childhood and adolescence. HNO
42:14-21

Balkany TJ, Finkel RS (1986) The dizzy child. Ear Hear 7:138-142
Baloh RW, Konrad HR, Hornrubia V (1975) Vestibulo-ocular function in patients with cerebellar atrophy. Neurology 25:160-168
Baloh RW, Jacobson K, Fife T (1994) Familial vestibulopathy: a
new dominantly inherited syndrome. Neurology 44:20-25
Barabas G, Matthews WS, Ferrari M (1983) Childhood migraine
and motion sickness. Pediatrics 72:188-190
Barr B, Lundstrom R (1961) Deafness following matern al rubella.
Acta Otolaryngol (Stockh) 53:413
Basser LS (1964) Benign paroxysmal vertigo of childhood. (A
variety of vestibular neuronitis.) Brain 87: 141-1 52

380
Beddoe GM (1977) Vertigo in childhood. Otolaryngol Clin North
Am 10:139-144
Belenky WM, Madgy DN, Leider JS, Becker q, Hotaling AJ (1993)
The enlarged vestibular aqueduct syndrome (EVA syndrome).
Ear Nose Throat J 72:746-751
Bickerstaff ER (1961) Basilar artery migraine. Lancet I: 15-17
Bixenman WW (1983) Congenital hereditary downbeat nystagmus. Can J OphthalmoI18:344-348
Blayney WA, Colman BH (1984) Dizziness in childhood. Clin
OtolaryngoI9:77-85
Bower CM, Cotton RT (1995) The spectrum of vertigo in children.
Arch Otolaryngol Head Neck Surg 121:9ll-915
Brandt Th, Bchele W (1984) Kindliche Schwindelformen. In:
Mortier W (ed) Moderne Diagnostik und Therapie bei
Kindern. Grosse, Berlin, pp 70-85
Brandt Th, Wenzel D, Dichgans J (1976) Die Entwicklung der
visuellen Stabilisation des aufrechten Standes beim Kind. Ein
Reifezeichen in der Kinderneurologie. Arch Psychiat Nervenkr
223:1-13
Bratt HD, Menelaus MB (1992) Benign paroxysmal torticollis of
infancy. J Bone Joint Surg Br 74:449-451
Britton BH, Block LD (1988) Vertigo in the pediatric and adolescent age group. Laryngoscope 96:139-146
Brunt D, Broadhead GD (1982) Motor proficiency traits of deaf
children. Res Q Exercise Sport 53:236-238
Butterfieid SA (1986) Gross motor profiles of deaf children.
Percept Mot Skills 62:68-72
Cassandro E, Mosca F, Sequino L, De Faleo FA, Campanella G
(1986) Otoneurological findings in Friedreich's ataxia and other
inherited neuropathies. Audiology 25:84-91
Chambers BR, Mai M, Barber HO (1985) Bilateral vestibular loss,
oscillopsia, and the cervico-ocular reflex. Otolaryngol Head
Neck Surg 93:403-407
Chinn HI, Smith PK (1955) Motion sickness. Pharmacol Rev 7:
33-83
Cohen HA, Nussinovitch M, Ashkenasi A, Straussberg R,
Kauschanksy A, Frydman M (1993) Benign paroxysmal torticollis in infancy. Pediatr Neurol 9:488- 490
Crook JP (1967) Congenital fistula in the stapedial footplate.
South Med J 60:ll68-ll70
Eil J, Prasher D, Rudge P (1984) Neuro-otological abnormalities in
Friedreich's ataxia. J Neurol Neurosurg Psychiatry 47:26-32
Enbom H, Magnusson M, Pyykk I (1991) Postural compensation
in children with congenital or early acquired bilateral vestibular loss. Ann Otol Rhinol LaryngoI100:472-478
Eviatar L, Eviatar A (1977) Vertigo in children: differential diagnosis and treatment. Pediatrics 59:833-838
Eviatar L (1994) Dizziness in children. Otolaryngol CI in North Am
27:557-571
Fried MP (1980) The evaluation of dizziness in children.
Laryngoscope 90: 1548-1560
Furman JMR, Baloh RW, Yee RD (1986) Eye movement abnormalities in a family with cerebellar vermian atrophy. Acta
Otolaryngol (Stockh) 101:371-377
Furman JMR, Wall C III, Pang D (1990) Vestibular function in
periodic alternating nystagmus. Brain 113:1425-1439
Gardner WJ, Frazier CH (1930) Bilateral acoustic neurofibromas: a
clinical study and field survey of a family of five generations
with bilateral deafness in thirty-eight members. Arch Neurol
Psychiat 23:266-302
Harrison MS (1962) Vertigo in childhood. J Laryngol Otol 76:
601-616
Hausler R, Toupet M, Guidetti G, Basseres F, Montandon P (1987)
Meniere's disease in children.Am J OtolaryngoI8:187-193
Hili J, Elliott C, Coloquhoun I (1992) Audiological, vestibular and
radiological abnormalities in Kallman's syndrome. J Laryngol
Otoll06:530-534
Horak FB, Shumway-Cook A, Crowe TK, Black FO (1988)
Vestibular function and motor proficiency of children with

Vertigo
impaired hearing or with learning disability and motor impairments. Dev Med Child NeuroI30:64-79
Horton WA, Wong V, Eldridge R (1976) Von Hippel-Lindau disease: Clinical and pathological manifestations in nine families
with 50 affected members. Arch Intern Med 136:769-777
Hoyt CS, Gelbart SS (1984) Vertical nystagmus in infants with
congenital ocular abnormalities. Ophthalmol Paediat Genet
(Amsterdam) 4:155-162
Huygen PLM, Verhagen WIM (1994) Peripheral vestibular and
vestibulo-cochlear dysfunction in hereditary disorders. J Vestib
Res 4:81-104
Kattah JC, Kolsky MP, Guy J, O'Doherty D (1983) Primary position
vertical nystagmus and cerebellar ataxia. Arch Neurol
40:310-314
Komune S, Nogami K, Inoue H, Uemura T (1993) Bilateral
Mondini dysplasia with normal hearing. ORL J
Otorhinolaryngol Relat Spec 55:143-146
Kumar A, Fishman G Torok N (1984) Vestibular and auditory
function in Usher's syndrome. Ann Otol Rhinol Laryngol
93:600-608
Lanzi G, Balottin U, Fazzi E, Tagliasacchi M, Manfrin M, Mira E
(1994) Benign paroxysmal vertigo of childhood: a long term
follow-up. Cephalalgia 14:458-460
Lenard HG, Voit T, Lamprecht A, Kahn T, Neuen-Jacob E,
Ruitenbeek W (1992) Sudden loss of hearing and vestibular
function, muscular weakness, and multiple white matter lesions
in preschool children. Neuropediatry 23:221-224
Longridge NS (1989) Progressive vestibular failure in childhood.
Acta Otolaryngol (Stockh) 468: 375-377
Mangabeira-Albernaz PL, Ganancam M (1988) Sudden vertigo of
central origin. Acta Otolaryngol (Stockh) 105:564-569
Murofushi T, Ouvrier RA, Parker GD, Graham RI, Da Silva M,
Halmagyi GM (1997) Vestibular abnormalities in CHARGE
association. Ann Otol Rhinol Laryngol 106: 129-134
Oosterveld WJ, Rademakers WJAC (1979) Nystagmus alternans.
Acta Otolaryngol (Stockh) 87:404-409
Pappas DG (1983) Hearing impairments and vestibular abnormalities among children with subclinical cytomegalovirus. Ann
Otol Rhinol Laryngol 92:552-557
Petroff MA, Simmons FB, Winzelberg J (1986) Two emerging perilymph fistula "syndromes" in children. Laryngoscope 96:
498-501
Pyykk I, Vesikivi M, Ishizaki H, Magnusson M, Juhola M (1991)
Postural control in blinds and in Usher's syndrome. Acta
Otolaryngol (Stockh) SuppI481:603-606
Rasmussen N, Johnsen NJ, Bohr VA (1991) Otologie sequelae after
pneumococcal meningitis: a survey of 164 consecutive cases
with a follow-up of 94 surviors. Laryngoscope 101:876-882
Rice WJ, Waggoner LG (1967) Congenital cerebrospinal fluid
otorrhea via a defect in the stapes footplate. Laryngoscope 77:
341-349
Sade J, Yaniv E (1984) Meniere's disease in infants. Acta
Otolaryngol (Stockh) 97:33-37
Savundra P, Skacel P, Rudge P (1996) Peripheral vestibulopathy in
sickle cell disease. J Audiol Med 5:61-66
Schlessel DA, Nedzelski JM (1992) Presentation oflarge vestibular
aqueduct syndrome to a dizziness unit. J Otolaryngol
21:265-269
Shibasaki H,Amashita Y, Motomura S (1978) Suppression of congenital nystagmus. J Neurol Neurosurg Psychiatry 41:
1078-1983
Shirabe S (1988) Vestibular neuronitis in childhood. Acta
Otolaryngol (Stockh) SuppI458:120-122
Simmons FB (1973) Patients with bilateral loss of caloric
response. Ann Otol 82: 175-178
Tahara T, Sekitani T, Imate Y, Kanesada K, Okami M (1993)
Vestibular neuronitis in children. Acta Otolaryngol (Stockh)
Suppl 503:49-52
Theunissen EJJM, Huygen PLM, Verhagen WIM (1989) Familial

Vertigo in childhood
vestibulocerebellar dysfunction: a new syndrome? J Neurol Sei
89:149-155
Tong KA, Harnsberger HR, Dahlen RT, Carey JC, Ward K (1997)
Large vestibular aqueduct syndrome: a genetic disease? Am J
RoentgenoI168:1097-1101
Trop I, Schloss MD, Polomeno R, Der Kaloustian V (1995) Usher
syndrome in four siblings from a consanguineous family of
Pakistani origin. J Otolaryngol 24: 102-104
Tusa RJ, Saada Jr AA, Niparko JK (1994) Dizziness in childhood. J
Child NeuroI9:261-274
Vanniasegaram I, Bellman S (1994) Vestibular function in
Stickler's syndrome. J Audiol Med 3:129-150
Verhagen WIM, Huygen PLM, Horstink NWIM (1987) Familial
congenital vestibular arefiexia. J Neurol Neurosurg Psychiatry
50:933-935
Verhagen WIM, Huygen PLM (1994) Central vestibular, vestibulo-

381
acoustic and oculomotor dysfunction in hereditary disorders. J
Vestib Res 4: 105-135
Verhagen WIM, Huygen PLM, Arts WFM (1996) Multi-system
signs and symptoms in X-linked ataxia carriers. J Neurol Sci
140:85-90
Vollertsen RS, McDonald TJ, Younge BR, Banks PM, Stanson AW,
Ilstrup DM (1986) Cogan's syndrome: 18 cases and a review of
the literature. Mayo Clin Proc 61:344-361
Weber PC, Perez BA, Bluestone CD (1993) Congenital perilymphatic fistula and associated middle ear abnormalities.
Laryngoscope 103:160-164
Weider DJ, Musiek FEC (1984) Bilateral congenital oval window
microfistula in a mother and son. Laryngoscope 94: 1455-1458
Weissman BM, Dell'Osso LF, Discenna A (1988) Downbeat nystagmus in an infant: Spontaneous resolution during infancy.
Neuro-Ophthalmology 8:317 - 319

SECTION H
Vertigo, dizziness, and falls
in the elderly

Vertigo, dizziness, and falls in the elderly

Basically most of the vestibular or non-vestibular


disorders that cause dizziness or vertigo also occur
in the elderly, some less frequently (e.g. basilar
migraine or motion sickness), others more frequently
(e.g. benign paroxy mal positioning vertigo or vascular vertigo). One might quest ion whether this
topic still warrants aseparate chapter.
There are several aspects that make management
in this age group special:
1. The frequency of dizziness and balance prob-

lems increases with advancing age to become the


most common presenting complaints in patients
older than 75.
2. Ageing of multisensory motor control, lack of
physical exercise, medication, and multimorbidity
in advanced age may all contribute to multifactorial conditions that present as dizziness or
postural imbalance.
3. Several abnormal clinical patterns of gait have
been described as typically manifesting in old
age, such as senile gait, frontal gait, or gait ignition failure.
4. The risk of falling rises with increasing age, and
falls in the elderly constitute an important health
problem.
The following points will be made in the discussion
of this topic:

Ageing is not a disorder but physiological senescence, which by itself (without additional
pathology) rarely causes dizziness or falls.
Dizziness and postural imbalance in the elderly
are frequently caused by vestibular, other sensory,
or central nervous system disorders.
Typical features of abnormal gait indicate different pathologies and the involvement of different
central nervous system structures.
Various conditions cause falls in the elderly;
so me can be successfuUy treated.

Physiological ageing of the vestibular


system
Sensory vestibular hair cells and central vestibular
neurons are differentiated cells that undergo continuing attrition from birth to old age, a senescence
that varies anatomically by region (Creasey and
Rapaport 1985; Vernadakis 1985). Ceilloss, electrophysiologic alterations, and changes in synapse
morphology as weil as in the supporting microenvironment have all been observed in portions of
the vestibular system (saccule, utricle, semicircular
canals, Scarpa's ganglia, and cerebellar vermis) of
aged animals and humans (Sloane et al. 1989a).
Progressive hair-cell reduction and loss of vestibular
nerve fibres and neurons in Scarpa's ganglion begin
at about age 40 to 50 (Johnsson 1971; Rosenhall
1973; Bergstrom 1973; Engstrom et al. 1974; Richter
1980). Cellular changes include shrinkage, alterations of stereocilia and kinocilia, appearance of
inclusion bodies, vesicle formation, degeneration of
otoconia (Engstrom et al. 1974; Richter 1980; Anniko
1983; Park et al. 1987; Gleeson and Felix 1987), and
synaptic pathology (Vernadakis 1985; Glick and
Bondareff 1979). Quanitative vestibular testing in
humans has shown alterations with age as weil, but a
decline with ageing is not a prominent feature of all
parameters, and many of these reported studies are
methodologically flawed (Sloane et al. 1989a).

Age-related changes in eye movements and


vestibulo-ocular reflexes
The senescence of human vestibulo-ocular reflexes
and optokinetic reflexes involves an amplitudedependent decrease in vestibulo-ocular reflex gain, a
shorter dominant vestibulo-ocular reflex time constant, an increase in phase lead, a decreased ability to

385

386

enhance or suppress the vestibulo-ocular reflex by


means of vision, and a lower optokinetic nystagmus
slow phase velocity saturation (Ura et al. 1991; Baloh
et al. 1993; Paige 1991,1994). The magnitude of these
changes is small relative to the variability of the data
(Peterka et al. 1990a; Paige 1994). High frequency
vestibulo-ocular reflex with voluntary head rotations
showed more inaccuracy in magnitude and timing
of the vestibulo-ocular reflex, and many of the aged
subjects were unable to generate voluntary fast head
movements (Hirvonen et al. 1997). Caloric test parameters showed no consistent age-dependent trend
(Peterka et al. 1990b). Older subjects show significant
reductions in initial pursuit acceleration (smooth
pursuit initiation to step-ramp stimuli) before saccades and in post-saccadic and peak pursuit velocity
(Morrow and Sharpe 1993). Elderly subjects also
generate less accurate saccades to moving targets,
but their drive for curiosity, as shown in another
study that measured exploratory eye movements,
can be well preserved (Daffner et al. 1994). Thus,
age-related changes in visual-vestibulo-ocular
responses are measurable, but they seem to vary and
do not quantitatively reflect the consistently and
progressively degenerative loss of vestibular hair
cells and nerve fibres.

Age-related changes in postural swayand


balance
Maintenance of postural balance is a multisensory
and sensorimotor ability that involves biomechanics,
sensory function, sensorimotor integration, and preprogramming of adjustments by exercise. There is
ample evidence of deterioration in many sensorimotor
systems underlying postural control, even in elderly
populations without obvious signs of disease (Horak
et al. 1989). Postural sway in static (Colledge et al.
1994b) and dynamic posturography (Camicioli et al.
1997; Wolfson et al. 1992) correlates with age and
functional parameters of balance (Tinetti Balance
Scale Score, timed one-leg standing), and there is an
increased heterogeneity of postural control abilities
in healthy older adults (Collins et al. 1995). Postural
responses to sudden translations of the support surface upon which the subject stands show small
increases in both EMG latencies and the time to
re ach the peak amplitude of centre of press ure
responses with increasing age (Peterka and Black
1990b), but the age-related trends are small relative
to the variability within the population. Studies on
age-related changes in open-loop and closed-loop
postural control mechanisms found that on the average, the elderly exhibit a greater delay before closedloop feedback mechanisms are called into play
(Collins et al. 1995). It has been repeatedly shown

Vertigo

that the sensorial weight of vision for postural control increases with increasing age, which makes an
elderly person particularly susceptible to altered
visual cues (Norre et al. 1987; Pyykk et al. 1990;
Peterka and Black 1990a). In contrast, children seem
to be more sensitive to alterations of somatosensory
cues (pressoreceptor and proprioceptor) (Peterka
and Black 1990a; Hytnen et al. 1993). Perturbation
of muscle spin dIes by vibration did not increase postural instability of elderly subjects (Pyykk et al.
1990). In a subsequent study postural control was
evaluated by vibration-induced body sway, measured
on a foam platform, and vibration sensation was
tested with a tuning fork (Kristinsdottir et al. 1997).
Vibration perception was the major determinant for
the magnitude of body sway. All senses show agerelated decrements, but the consequences for postural
control differ. Whereas the loss of proprioception in
the legs may contribute to increased body sway in
the elderly, it is the visual input, despite impaired
visual function, which plays a dominant role in old
age, not only for postural control but also for spatial
orientation and perception of self-motion (Paige
1994).
Abnormal central processing of sensory input
may compromise balance in the setting of postural
perturbations to a greater degree in the advanced
elderly, who under experimental conditions, e.g.
showed diminished adaptation to repeated platform
rotations (Camicioli et al. 1997). Effective balance
training studies have employed the principle of overload to elicit improvements in stability not only in
young adults (Brandt et al. 1981), but also in older
adults (Seidler and Martin 1997). Balance training
consisting of exercises designed to stress balance
and co ordination was performed three times a week
for 5 weeks; however, measurements of postural
sway showed that this was only moderately effective
for training elderly adults (Seidler and Martin 1997).
In addition to the physiological ageing processes,
the occurrence of pathological conditions of many
forms, even in early, subclinical stages, contributes to
any disequilibrium present in an elderly individual.
Even though different pathological processes will
affect different functional components of postural
control, the combined effects of age and pathology may
appear on casual inspection to be a non-specific generalised increase in disequilibrium (Horak et al. 1989).

Cautious senile gait and /lhighest


level gait disorders"
Walking is one of the most common of all human
movements; it is learned during the first year of life

387

Vertigo, dizziness, and falls in the elderly

and perfected by around age 7. After age 60, this ability starts to decline, and the elderly gradually slow
down (Prince et al. 1997). A population-based study
showed that 15% of the subjects over age 60 had
some abnormality of gait (Newman et al. 1960), and
postural imbalance and gait dis orders significantly
contributed to the risk of falling (Tinetti et al. 1988).
The so-called senile gait is more appropriately called
cautious gait. Many older people adopt it to a greater
or lesser degree to compensate for arthritis, pain,
sensory or vestibular impairment, or simply from
fe ar of falling (Marsden and Thompson 1996). When
walking, elderly persons responded to a probe auditory reaction time task with greater delays than
young adults. This suggests that the elderly require a
greater proportion of attentional resources to be
allocated to the demands of balance (Lajoie et al.
1996).
"Anyone whose balance is insecure, for whatever
reason, will attempt to compensate. Normal compensation is best illustrated by our natural re action to
walking on ice. The feet are placed apart to widen
the base; the body, hips, and knees are bent to place
the centre of gravity firmly over the widened base;
the arms are held somewhat abducted and ftexed in
anticipation of unexpected threats to balance.
Locomotion proceeds with small steps on this wide
base with a ftexed posture. We walk the same way on
a rolling deck of a ship" (Mars den and Thompson
1996).
Using analysis of covariants, Elble and co-workers
(1992) found that these changes in gait were distinctly attributable to the reduced stride of older
people and not to age per se; they suggest that many
gait disturbances in elderly people are similar,
regardless of aetiology, because the characteristics of
these gait disturbances are heavily veiled by nonspecific stride-dependent changes that comprise the
syndrome of senile gait. It was also speculated that
changes in gait pattern with increasing age are associated with decreasing muscle strength (Nigg et al.
1994).
Nevertheless, as clinicians we feel that we are able
to differentiate the cautious gait of the elderly from
other "highest level gait dis orders" (Nutt et al. 1993)
such as frontal disequilibrium, isolated gait ignition
failure (Atchison et al. 1993), or frontal gait disorder
(see Table 27.1 for classification and Table 27.2 for
terminology). Subcortical disequilibrium has been
reported to be an acute consequence of thalamic
basal ganglia or midbrain stroke (see thalamic astasia,
p. 192). In frontal disequilibrium there appears to be
a breakdown of the organisation of leg movements
required for locomotion, so that the gait is often
bizarre and the feet frequently cross or move in the
wrong direction (Marsden and Thompson 1996).

Table 27.1. Classifieation of gait syndromes


I Lowest-Ievel goit disorders
A. Peripheral skeletomuseular problems
Arthritie gait
Myopathie gait
Peripheral neuropathie gait
B. Peripheral sensory problems
Sensory ataxie gait
Vestibular ataxie gait
Visual ataxie gait

IJ Middle-Ievel gait disorders


Hemiplegie gait
Paraplegie gait
Cerebellar ataxie gait
Parkinsonian gait
Choreie gait
Dystonie gait
111 Highest-Ievel gait disorders
Cautious gait
Subeortieal disequilibrium
Frontal disequilibrium
Isolated gait ignition failure
Frontal gait disorder
Psyehogenie gait disorder
From Nutt et al. (1993).

Table 27.2. Terminology used to deseibe the elinieal patterns of gait in


this chapter compared with those in previous publieations
Proposed terminology Previous terms

Lesions

Cautious

Elderly gait
Senile gait

Musculoskeletal
Peripheral nervous
lesions
Central nervous lesions

Subcortical
disequilibrium

Tottering
Astasia-abasia
Thalamie astasia

Midbrain
Basal ganglia
Thalamus

Frontal disequilibrium Gait apraxia


Frontal ataxia
Astasia-abasia

Frontal lobe and white


matter connections

Isolated gait ignition


failure

Gait apraxia
Magnetic gait
Slipping cluteh gait
Lower half
parkinsonism
Arteriosclerotic
parkinsonism
Trepidant abasia
(Petren's gait)

Frontal lobe, white


matter connections, and
basal ganglia

Frontal gait disorder

Marche apetits pas


Frontal lobe and white
Magnetie gait apraxia matter lesions
Arteriosclerotic
parkinsonism
Parkinsonian ataxia
Lower half parkinsonism
Lower body parkinsonism

From Nutt et al. (1993).

Vertigo

388

Frontal disequilibrium is also described as gait


apraxia (e.g. in normal pressure hydrocephalus),
frontal ataxia, or astasia-abasia, which indicates dysfunction of the frontal lobe and white matter connections. Isolated gait ignition failure describes an
inability to initiate and sustain locomotion with
start-and-turn hesitation, shuffting and freezing, but
relatively normal gait once locomotion is initiated;
the posture is upright, there is good arm swing, a
normal stride length and no festination (Atchison et
al. 1993). This makes it different from posture and
gait in Parkinson's syndrome (Dietz et al. 1988;
Sudarsky 1990; Nutt et al. 1993). The so-called frontal
gait disorder shares some features of frontal disequilibrium and gait ignition failure; it is characterised
by a variable base (narrow-to-wide), difficulty starting to walk, short steps, shuffting, hesitation on turns
with freezing and moderate disequilibrium.
The crucial questions about these syndromes of
highest level gait disorders are to what extent are
they separate entities and how much overlap is there
among them (Elble et al. 1992; Marsden and
Thompson 1996). For differentiation from psychogenic dis orders of stance and gait, see Chap. 31
(p.461).

Falls in the elderly


The risk of falls increases with increasing age. In
studies on the frequency of falls in randomly selected
community populations, annual prevalence rates
were about 30% for those older than age 65
(Campbell et al. 1981; Prudham and Evans 1981),
35% for those older than age 70 (Campbell et al.
1989), and about 40% for those older than age 75
(Tinetti et al. 1988; Downton and Andrews 1991).
Healthy elderly subjects are less likely to fall (GabeIl
et al. 1985), but 60-70% of those who have fallen in
the previous year are likely to fall during the following 12 months (Nevitt et al. 1989; Downton 1996). f
those residing in nursing hornes, 45-75% have been
reported to fall at least once a year (Gryfe et al. 1977;
Tinetti 1987), but the average rate offalls in a nursing horne has also been estimated to be two falls per
patient per year (Baker and Harvey 1985). Accidental
injury is the sixth leading cause of death among the
elderly, and a majority of these injuries result from
falls (Baker and Harvey 1985). In addition to their
substantial adverse economic effects, falls and fallrelated injuries (primarily hip fractures) have major
medical and social consequences (Thapa and Ray
1996), including loss of ambulation (Cummings et al.
1985), the early move to a nursing horne (Ray et al.

1987) and increased mortality (Ray et al. 1990). The


mortality rate of old people who have fallen is
greater than that of non-fallers (Wild et al. 1980),
and falls are one of several factors found to be significant predictors of mortality in a community population (Campbell et al. 1985).
Ageing, lack of physical exercise, medications,
multiple cardiovascular and sensorimotor disorders,
as weIl as behavioural and environmental factors are
involved in the aetiology of falls in the elderly
(Sudarsky 1990; Downton 1994,1996; Thapa and Ray
1996). Different studies identify different major
causes of falls. For example, Jntti et al. (1993)
reported that the most important factor associated
with falls is impaired visual acuity. In addition, more
than 50% of the fallers suffered from dementia.
About 90% of falls are not caused by or associated
with loss of consciousness (Prudham and Evans
1981; Downton 1996), and of 500 falls in 202 people
aged 50-90 years, 125 drop attacks occurred in 58
individuals (Table 27.3; Sheldon 1960).
Lee and Marsden (1996) define a drop attack as "a
sudden fall, with or without loss of consciousness,
due either to collapse of postural muscle tone (a
negative phenomenon) or abnormal muscle contraction in the legs (a positive phenomenon)". The multiple symptomatic causes are summarised in Table
27.4. Two-thirds of patients with drop attacks have
no obvious cause (Table 27.5) - these are called idiopathic drop attacks - and the vast majority of those
suffer only minor injury and have a favourable prognosis (Meissner et al. 1986; Lee and Marsden 1996).
For vestibular drop attacks see Chap. 5 (p. 94).
Vasovagal syncope (Kapoor 1994) and orthostatic
hypotension may cause a sudden brief loss of consciousness due to transient global impairment of
cerebral circulation with dizziness followed by
Table 27.3. Falls in old age
500 falls in 202 people aged 50-90 years
Accidental falls
Drop attacks
Trips
Vertigo
CNS lesions
Extension of neck
Postural syncope
Leg weakness
Falls out of bed or chair
Uncertain

171
125

59 falls resulted in a fractured femur


Drop attacks (125 of the 500 falls) occurred in 58 individuals.
26 of the 58 patients sustained fractures or dislocations.
51 ofthe 58 patients had more than 1 fall.
From Sheldon (1960).

53
37
27
20
18
16
10
23

389

Vertigo, dizziness, and falls in the elderly


Table 27.4. Causes of drop attacks

ISymptomotic drop ottocks


1.

2.

3.
4.

5.
6.

7.

8.

9.
10.

11.

Drop attacks associated with weak legs:


Muscular dystrophy
Inclusion body myositis
Myasthenia gravis
Poly myositis
Neurogenie atrophy
Intermittent spinal ischaemia
Neurodegenerative disease:
Idiopathie Parkinson's disease
Multiple system atrophy
Huntington's disease
Progressive supranuclear palsy
Corticobasal degeneration
Alzheimer's disease
Syncope, cardiac dysrhythmia, cyanotie heart disease
Transient ischaemic attacks:
Vertebrobasilar ischaemia
Extrinsie compression: cervical spondylosis
Intrinsie obstruction
Bilateral anterior cerebra I artery ischaemia
Peripheral vestibular disease:
Meniere's syndrome
Epilepsy:
Lennox-Gastaut syndrome
Epilepsy with myoclonic-astatic seizures
Juvenile myoclonie epilepsy
Complex partial frontal seizures
Myoclonie jerks; positive and negative:
Post-hypoxie myoclonus
Progressive myoclonie epilepsy or ataxia:
Unverricht-Lundborg disease
Lafora body disease
Sialidosis
Ceroid lipofucinosis
Mitochondrial disease
Spinocerebellar degenerations
GM2 gangliosidosis
Hallervorden-Spatz disease
Juvenile neuroaxonal dystrophy
Startle reactions:
Psychogenie startle reactions:
Exaggerated normal physiologie reaction
Hyperekplexia:
Idiopathic
Familial: autosomal dominant
Symptomatic:
Post-hypoxie encephalopathy
Post-traumatie encephalopathy
Viral encephalitis
Multiple sclerosis
Paraneoplastic syndrome
Cerebral abscess
Brainstem stroke
Other drop attacks which can be precipitated by sudden emotional changes:
Cataplexy
Paroxysmal kinesigenie choreoathetosis
Focal structural central nervous system lesions:
Frontal lobe lesions:
Tumour
Scar
Lower brainstem, upper cervieal cord, and/or cerebellar compression:
Odontoid process fracture
Arnold-Chiari malformation
Posterior fossa arachnoid cyst
Metastatic cerebellar carcinoma
Cerebellar vermis tumour
Hydrocephalus or periventrieular white matter pathology:
Normal pressure hydrocephalus
Obstructive hydrocephalus associated with
Third ventricle colloid cyst
Third ventricle meningioma
Fourth ventricle ependymoma
Congenital obstruction of aqueduct ofSylvius
Binswanger disease

IIldiopothic drop attocks


From Lee and Marsden (1996).

Table 27.5.

Natural history of drop attacks

166 patients seen at Mayo Clinic 1976-1983.


58 excluded, rnost because of cardiogenic causes with changes in
consciousness
108 (76 women and 32 men) met definition of"A falling speil occurring
without warning or postictal symptoms, with immediate righting, and without
loss of awareness or consciousness."
Mean age 70 years (most over the age of 60); follow-up 6-7 years
Causes
Cardiac
Cerebrovascular disease
Combined cardiac and cerebrovascular disease
Seizures
Vestibular disease
Psychogenic
Unknown

12%
7%
8%
5%
3%
1%
64%

Unknown causes (69 ca ses) included


Hypertension
Cardiac disease
Migraine
Diabetes
Alcohol
Cataplexy
Cervical spondylosis

41 %
25%
25%
6%
3%
1%
19%

92 of the 108 were alive at follow-up.


84% of 58 untreated patients were asymptomatic.
Stroke rate (0.5% per year), similar to controls.
From Meissner et al. (1986).

blackout and falls (Downton 1996). Injuries and fractures are more likely to occur with syncopal falls
than in falls without loss of consciousness.
Medications significantly increase the risk of falls
due to sedation as weIl as psychomotor and automotor impairment. The elderly are particularly susceptible to these effects (Table 27.6; Thapa and Ray
1996).

Dizziness in the elderly


Community surveys show dizziness affects more
than 30% of people age 65 and over (Sloane et al.
1989b; Colledge et al. 1994a; Grimby and Rosenhall
1995). In office practice, the incidence of dizziness
increases with patient age, becoming the most common presenting symptom of those age 75 and older
(Sloane et al. 1989a). Since dizziness in older people
can represent many different abnormal sensations,
the initial task of the examining physician is to
determine what the patient means by dizziness: presyncopallight-headedness, postural disequilibrium,
or vertigo (Baloh 1992). Presyncopallight-headedness, or the sensation of an impending faint may be
associated with the feeling of unsteadiness. Postural
imbalance or unsteadiness unrelated to an abnormal

390

Vertigo

Table 27.6. Use in two elderly populations of medications hypothesised to increase risk of falling'
Medication

- - - - - - -

Diuretics
Antihypertensives
Psychotropic drugs
Benzodiazepines
Long half-life
Short half-life
Other hypnotics e
Antidepressants d
Antipsychotics

Effect

----

Prevalence of use (%)b

----

--

Postural hypotension
Postural hypotension, sedation
Sedation, postural hypotension
Sedation
Sedation
Sedation
Sedation, postural hypotension
Sedation, postural hypotension

Saskatchewan 1985
(n= 3228)

Tennessee Medicaid 1986


(n = 4350)

47

56

40
34

50

10
13

15

6
8

10
4

21

, Data obtained from a stratified random sampie drawn so that the age, sex, race, and nursing horne status of the sampie would be identical to that of
elderly hip fracture patients.
b Includes any prescription filled during the one-year period.
eincIudes diphenhydramine, chloralhydrate, hydroxyzine, and meprobamate.
d Excludes monoamine oxidase inhibitors.
From Ray and Griffin (1990).

head sensation may also be described as dizziness,


whereas vertigo means the apparent sensation of
rotation or tilt.
Analogously to the frequency of forms of vertigo
reported to occur in children, many findings in the
elderly seem to depend on the setting of the study. In
a retrospective study of more than 1000 cases of
patients age 70 or older, who presented to a dizziness
dinie, 40% had benign paroxysmal positioning vertigo, followed by other vestibular or neurological
dis orders such as Meniere's disease, vestibular
neuritis, cerebral vascular episodes, and tumours
(Katasarkas 1994). In contrast, areport from a geriatrie medicine unit found that benign paroxysmal
positioning vertigo accounted for only 4% of cases of
dizziness, whereas cerebrovascular dis orders, cervical spondylosis (?), anxiety, and hyperventilation
were indicated to be major causes of dizziness in
older people (Colledge et al. 1996).
Dizziness in elderly patients is a diagnostie challenge because of multimorbidity. Of 740 patients
older than 65 who reported dizziness, a specific
cause was determined in only 21 % (Belal and Glorig
1986); the remaining 79% were diagnosed to have
"presbyastasis", a term signifying disequilibrium of
ageing. This figure is unsatisfying. Certainly careful
otoneurological and neuro-ophthalmological evaluation will disdose a higher percentage of identifiable
causes in the future. A considerable number of
vestibular dis orders can be seen in elderly patients,
particularly benign paroxysmal positioning vertigo
(Bioom and Katasarkas 1989), bilateral or asymmetrie
vestibular deficits, Meniere's disease, or central
vestibular disorders. Careful dinieal reevaluation of
26 patients over age 75 who complained of disequi-

librium and in whom no cause was evident after the


first dinical evaluation revealed that seven patients
had profoundly reduced vestibular function; and the
remaining 19 patients had a significantly lower average vestibular function than controls (Fife and Baloh
1993).
Thus, dizziness in older patients should prompt
the physician to carefully test peripheral vestibular
function (for bilateral vestibular failure, see Chap. 8)
and benign paroxysmal positioning vertigo (see
Chap. 16). Other more rare causes indude vestibular
paroxysmia (Chap. 7) or somatosensory loss in
polyneuropathy. The latter may become stronger
symptomatically, if a polyneuropathy combines with
visual impairment (p. 447). Gait dis orders in the
elderly may also present as subjective dizziness in
the initial stage. It is our own experience that
patients with Steele-Riehardson -Olszewski syndrome report dizziness or to-and-fro vertigo to be
the initial symptom of the condition, which is later
best diagnosed by the combination of a vertical
supranudear palsy with downward gaze abnormality
and postural instability with unexplained falls
(Litvan et al. 1996). Finally, depression in old age
may also present as dizziness and unsteadiness or
more frequently may exaggerate an organic dizziness, postural instability, or gait dis order.
In summary, diagnosis and management of vertigo and dizziness in the elderly is difficult in two
respects: first, to identify the multitude of possible
causes which may summate or potentiate their
effects, and second, to relate the particular dis order
to the presenting signs and symptoms. Studies are
available that try to differentiate between dysfunctions found in the vestibular, visual, or proprioceptive

Vertigo, dizziness, and falls in the elderly

systems and to determine the probability of the


principal cause for the presenting dizziness (Davis
1994). However, when a depressive patient with
Parkinson's syndrome and arte rial hypotension, sensory polyneuropathy, and impairment of vision
complains of dizziness, the physician is highly justified to be cautious about assigning a specific clinical
weight to the major cause of his complaints.
Management must concentrate on treating those
possible causes that can be treated and, equally
important, on motivating the patient to engage in
physical exercise and balance training rather than to
withdraw socially and physically or even immediately
res ort to gait aids.

References
Anniko M (1983) The aging vestibular hair cel!. Am J Otolaryngol
4:151-160
Atchison PR, Thompson PD, Frackowiak RSJ, Marsden CD (1993)
The syndrome of gait ignition failure: areport of six cases.
Movem Disord 8:285-292
Baker SP, Harvey AH (1985) Fall injuries in the elderly. Clin
Geriatr Med 1:501-508
Baloh RW (1992) Dizziness in older people. J Am Geriatr Soc
40:713-721
Baloh RW, Jacobson KM, Socotch TM (1993) The effect of aging
on visual-vestibuloocular responses. Exp Brain Res 95:509-516
Belal A, Glorig A (1986) Dysequilibrium of aging (presbyastasis). J
Laryngol Otol 100: 1037 -1041
Bergstrom B (1973) Morphology of the vestibular nerve H. The
number of myelinated vestibular nerve fibres in man at various
ages.Acta Otolaryngol (Stockh) 76:173-179
Bloom J, Katasarkas A (1989) Paroxysmal positional vertigo in the
elderly. J OtolaryngoI18:96-98
Brandt Th, Krafczyk S, Malsbenden I (1981) Postural imbalance
with head extension: improvement by training as a model for
ataxia therapy.Ann NY Acad Sei 374:646-649
Camicioli R, Panzer VP, Kaye J (1997) Balance in the healtby elderly.
Posturography and clinical assessment. Arch Neurol
54:976-981
Campbell AJ, Borrie MJ, Spears GF (1989) Risk factors for falls in a
community-based prospective study of people 70 years and
older. J Gerontol44:M 112-M 117
Campbell AJ, Reinken J, Allan BC, Martinez GS (1981) Falls in old
age: a study of frequency and related clinical factors. Age
Ageing 10:264-270
Campbell AJ, Diep C, Reinken J, McCosh L (1985) Factors predicting mortality in a total population sampie of elderly. J
Epidemiol Commun Health 39:337-342
Colledge NR, Wilson JA, Macintyre CCA, MacLennan WJ ( 1994a)
The prevalence and characteristics of dizziness in an elderly
community. Age Ageing 23: 117-120
Colledge NR, Cantley P, Peaston I, Brash H, Lewis S, Wilson JA
(1994b) Ageing and balance: the measurement of spontaneous
sway by posturography. Gerontology 40:273-278
Colledge NR, Barr-Hamilton RM, Lewis SI, Sellar RJ, Wilson JA
(1996) Evaluation of investigations to diagnose the cause of
dizziness in elderly people. Br Med J 313:788-792
Collins JJ, De Luca CJ, Burrows A, Lipsitz LA (1995) Age-related
changes in open-loop and closed-loop postural control
mechanisms. Exp Brain Res 104:480-492

391
Creasey H, Rapaport SI (1985) The aging human brain. Ann
NeuroI17:2-10
Cummings SR, Kelsey JL, Nevitt MC, O'Dowd KJ (1985)
Epidemiology of osteoporosis and osteoporotic fractures.
Epidemiol Rev 7:178-208
Daffner KR, Seinto LM, Weintraub S, Guinessey 1, Mesulam MM
(1994) The impact of ageing on curiosity as measured by
exploratory eye movements. Arch Neurol 51 :368-376
Davis LE (1994) Dizziness in elderly men. J Am Geriatr Soc
42:1184-1188
Dietz V, Berger W, Horstmann GA (1988) Posture in Parkinson's
disease: impairment of reflexes and programming. Ann Neurol
24:660-669
Downton JH (1994) Editorial: falls in the elderly-epidemiology,
classification and causes. J Audiol Med 3:3-13
Downton JH (1996) Falls in the elderly: a clinical view. In:
Bronstein AM, Brandt Th, Woollacott M (eds) Clinical dis orders
of balance, posture and gait. Edward Arnold, London, pp
326-341
Downton JH, Andrews K (1991) Prevalence, characteristics and
factors associated with falls among the elderly living at horne.
Aging 3:219-228
Elble RJ, Hughes L, Higgins C (1992) The syndrome of senile gait.
J NeuroI239:71-75
Engstrom H, Bergstrom B, Rosenhall U (1974) Vestibular sensory
epithelia!.Arch OtolaryngoI100:411-418
Fife TD, Baloh RW (1993) Disequilibrium of unknown cause in
older people. Ann Neurol 34:694-702
GabeIl A, Simons MA, Nayak USL (1985) Falls in the healtby elderly:
predisposing causes. Ergonomics 28:965-975
Gleeson M, Felix H (1987) A comparative study of the effect of age
on the human cochlear and vestibular neuroepithelia. Acta
Otolaryngol (Stockh) SuppI436:103-109
Glick R, Bondareff W (1979) Loss of synapses in the cerebellar
cortex of the senescent rat. J Gerontol 34:818-822
Gryfe JS, Amies A, Ashley MJ (1977) A longitudinal study of falls
in an elderly population: l. ineidence and morbidity. Age Ageing
6:201-210
Horak FB, Shupert CL, Mirka A (1989) Components of postural
dyscontrol in elderly: a review. Neurobiol Aging 10:727-738
Hirvonen TP, Aalto H, Pyykk I, Juhola M, Jntti PO (1997)
Changes in vestibulo-ocular reflex of elderly people. Acta
Otolaryngol (Stockh) SuppI529:108-11O
Hytnen M, Pyykk I, Aalto H, Starck J (1993) Postural control
and age. Acta Otolaryngol (Stockh) 113:119-122
Jntti PO, Pyykk VI, Hervonen ALJ (1993) Falls among elderly
nursing horne residents. Public Health 107:89-96
Johnsson L (1971) Degenerative changes and anomalies of the
vestibular system in man. Laryngoscope 81: 1682-1694
Kapoor WN (1994) Syncope in older persons. J Am Geriatr Soc
42:426-436
Katasarkas A (1994) Dizziness in aging: a retrospective study of
1194 cases. Otolaryngol Head Neck Surg 110:296-301
Kristinsdottir EK, Jarnlo GB, Magnusson M (1997) Aberrations in
postural contro!, vibration sensation and some vestibular findings in healthy 64-92 year-old subjects. Scand J Rehab Med
29:257-265
Lajoie Y, Teasdale N, Bard C, Fleury M (1996) Upright standing
and gait: are there changes in attentional requirements related
to normal aging? Exp Aging Res 22:185-198
Lee MS, Marsden CD (1996) Drop attacks. In: Bronstein AM,
Brandt Th, Woollacott M (eds) Clinical dis orders of balance,
posture and gait. Edward Arnold, London, pp 177-187
Litvan I, Agid Y, Jankovic J, Goetz C, Brandel JP, Lai EC, Wenning
G, D'Olhaberriague U, Verny M, Chaudhuri KR, McKee A,
Jellinger K, Bartko JJ, Magone CA, Pearce RK (1996) Accuracy of
clinical criteria for the diagnosis of progressive supranuclear
palsy (Steele-Richardson-Olszewski syndrome). Neurology
46:922-930

392
Marsden CD, Thompson PD (1996) Frontal gait disorders. In:
Bronstein AM, Brandt Th, Woollacott M (eds) Clinical dis orders
of balance, posture and gait. Edward Arnold, London, pp
188-193
Meissner I, Wiebers DO, Swanson JW, O'Falion WM (1986) The
natural his tory of drop attacks. Neurology 36:1029-1034
Morrow MJ, Sharpe JA (1993) Smooth pursuit initiation in young
and elderly subjects. Vision Res 33:203-210
Nevitt MC, Cummings SR, Kidd S, Black D (1989) Risk factors for
recurrent non -syncopal falls. A prospective study. JAMA
261 :2663-2668
Newman G, Dovenmuehle RH, Busse EW (1960) Alterations in
neurological status with age. J Am Geriatr Soc 8:915-917
Nigg BM, Fisher V, Ronsky JL (1994) Gait characteristics as a function of age and gender. Gait Posture 2:213-220
Norre ME, Forrez G, Beckers A (1987) Vestibular dysfunction
causing instability in aged patients. Acta Otolaryngol (Stockh)
104:50-55
Nutt JG, Marsden CD, Thompson PD (1993) Human walking and
high er-level gait dis orders, particularly in the elderly.
Neurology 43:268-279
Paige GD (1991) The ageing vestibulo-ocular reflex (VOR) and
adaptive plasticity. Acta Otolaryngol (Stockh) Suppl
481:297-300
Paige GD (1994) Senescence of human visual-vestibular interactions: smooth pursuit, optokinetic, and vestibular control of eye
movements with aging. Exp Brain Res 98:355-372
Park JC, Hubel SB, Woods AD (1987) Morphometric analysis and
fine structure of the vestibular epithelium of aged C57BL/6Nia
mice. Hear Res 28:87-96
Peterka RJ, Black FO (1990a) Age-related changes in human posture control: sensory organization tests. J Vestib Res 1:73-85
Peterka RJ, Black FO (1990b) Age-related changes in human posture control: motor coordination tests. J Vestib Res 1:87-96
Peterka RJ, Black FO, Schoenhoff MB (1990a) Age-related changes
in human vestibulo-ocular and optokinetic reflexes: pseudorandom rotation tests. J Vestib Res 1:61-71
Peterka RJ, Black FO, Schoenhoff MB (1990b) Age-related changes
in human vestibulo-ocular reflexes: sinusoidal rotation and
caloric tests. J Vestib Res 1:49-59
Prince F, Corriveau H, Hebert R, Winter DA (1997) Gait in the
elderly. Gait Posture 5: 128-135
Prudham D, Evans JG (1981) Factors associated with falls in the
elderly: a community study. Age Ageing 10: 141-146
Pyykk I, Jntti PO, Aalto H (1990) Postural control in elderly subjects.Age Ageing 19:215-221

Vertigo
Ray WA, Federspiel CF, Baugh DK, Dodds S (1987) Interstate variation in elderly Medicaid nursing horne populations. Med Care
25:738-752
Ray WA, Griffin MR (1990) Prescribed medications and the risk of
falling. Top Geriatr Rehab 5:12-20
Ray WA, Griffin MR, Baugh DK (1990) Mortality following hip
fracture before and after implementation of the prospective
payment system. Arch Intern Med 150:2109-2114
Richter E (1980) Quantitative study ofhuman Scarpa's ganglion
and vestibular sensory epithelia. Acta Otolaryngol (Stockh)
90:199-208
Rosenhall U (1973) Degenerative patterns in the aging human
vestibular neuroepithelia. Acta Otolaryngol (Stockh)
76:208-228
Seidler RD, Martin PE (1997) The effects of short term balance
training on the postural control of older adults. Gait Post ure
6:224-236
Sheldon JH (1960) On the natural history of falls in old age. Br
Med J II: 1685-1690
Sloane PD, Baloh RW, Honrubia V (1989a) The vestibular system
in the elderly: clinical implications. Am J Otolaryngol 10:
422-429
Sloane P, Blazer D, George LK (1989b) Dizziness in a community
elderly population. J Am Geriatr Soc 37:101-108
Sudarsky L (1990) Geriatrics: gait dis orders in the elderly. N Engl J
Med 322:1441-1446
Thapa PB, Ray WA(1996) Medications and falls and fall-related
injuries in the elderly. In: Bronstein AM, Brandt Th, Woollacott
M (eds) Clinical dis orders of balance, posture and gait. Edward
Arnold, London, pp 301-325
Tinetti ME (1987) Factors associated with serious injury during
falls by ambulatory nursing horne residents. J Am Geriatr Soc
35:644-648
Tinetti ME, Speechley M, Ginter SF (1988) Risk factors for falls
among elderly persons living in the community. N Engl J Med
319:1701-1707
Ura M, Pfaltz CR,Allum JHJ (1991) The effect of age on the visuoand vestibulo-ocular reflexes of elderly patients with vertigo.
Acta Otolaryngol (Stockh) SuppI481:399-402
Vernadakis A (1985) The aging brain. Clin Geriatr Med 1:61-94
Wild D, Nayak USL, Isaacs B (1980) Characteristics of old people
who fell at horne. J Clin Exp GerontoI2:271-278
Wolfson L, Whip pie R, Derby CA, Amerman P, Murphy T, Tobin
IN, Nashner L (1992) Adynamie posturography study of balance in healthy elderly. Neurology 42:2069-2075

SECTION I
Drugs and vertigo

Drugs and vertigo

The list of drugs that may have adverse effects on


hearing or balance is impressive (see also Chap. 24).
It includes anaesthetic ,anticonvulsants, antidepressants, analgesics, antidiabetics, antihypertensive
agents, anti-infiammatory drugs, contraceptives,
cytotoxic agents, cardiovascular drugs, sedatives,
and tranquillizers (Table 28.1; Ballantyne and
Ajodhia 1984; Black and Pesznecker 1993; Rascol et
al. 1995). Historically, the first identified ototoxic
drugs were naturally occurring substances, e.g. mercury as a primary treatment for syphilis, followed by
the first aminoglycoside antibiotic effective against
tuberculosis (Black and Pesznecker 1993).
Irreversible deafness and disturbance of balance
were reported in the first clinical trial of streptomycin (Hinshaw and Feldman 1945). An exhaustive
treatment of the extensive literature on drugs causing vestibular and oculomotor dysfunction is, however, beyond the scope of this monograph.
There is no common syndrome or mechanism of
drug-induced vertigo. The spectrum of its side
effects is broad. For example, beta-blockers may
induce bradycardia with unspecific drowsiness or
dizziness (Cruickshank 1981). More severe side
effects include dizziness with hypoglycaemia
(HarrilI1951; Currier 1970), positional alcohol nystagmus/vertigo (p. 286), cerebellar ataxia and oculomotor dysfunction in acute diphenylhydantoin
intoxication (Nozue et al. 1973), and bilateral
vestibu!ar loss with oscillopsia and unsteadiness of
gait due to gentamicin toxicity (Ramsden and AckriJl
1982). Drug-induced patterns of pathological eye
movements include positional, downbeat and gazeevoked nystagmus; they generally indicate transient
fiocculus dysfunction (Esser and Brandt 1983).
Thus, patients who present with vertigo and centra! ocular motor abnormalities should always be
questioned about the drugs they are taking. This is
of particular importance in the following conditions:

"unspecific dizziness" without associated neurological signs


fiuctuating cerebellar and/or oculomotor signs

fiuctuating dizziness and unsteadiness with


unusually prolonged daily sleep phases (>8 h)
repeated falls in the horne environment
positional vertigo and/or nystagmus
bilateral vestibular and/or hearing loss

Ototoxic agents
Aminoglycoside antibiotics can permanently damage the cochlea or the vestibular labyrinth (Wersll
1995; Murofushi et al. 1997). Whereas kanamycin,
neomycin and vancomycin are predominantly
cochleotoxic, gentamicin, streptomycin and
tobramycin are predominantly vestibulotoxic
(Ballantyne 1984). Loss ofvestibular function (Chap.
8) impairs perception of head motion and visual
stabilisation of gaze in space as a result of insufficient VOR. Caloric responses, per- and post-rotatory
nystagmus (bilateral vestibular fa ilure , p. 127), as
weH as optokinetie after-nystagmus are absent (Hain
and Zee 1991). Oscillopsia and unsteadiness of gait
are the major complaints of these patients. They also
have a considerable postural instability when visual
or somatosensory cues for maintaining balance are
alte red (Herdman et al. 1994).
Histologically, streptomycin causes degeneration
of the hair cells and ampullary cristae (Lindeman
1969; Schuknecht 1974) and circumscribed defects of
otoconial membrane (Jonsson et al. 1980).
Gentamicin also damages the secretory cells of the
inner ear, a side effect that is used in the treatment of
endolymphatic hydrops in Meniere's disease
(p. 83). Ototoxie effects are augmented by impaired
renal function (Ballantyne 1970), previous treatment
with diuretics (West et al. 1973), or antimalarial
drugs (Nilges and Northern 1971). The fact that
enzymatic transformation of gentamicin to a
metabolite or an "activated molecule" may precede
toxic actions, demonstrates that its capability for

395

396

Vertigo

Table 28.1. Drugs causing vertigo (modified and shortened after Ballantyne and Ajodhia 1984)

Drugs aeting on the eardiovaseular system

1. Antihypertenslve drugs
Alkaloids of veratrum viride + epithiazide, bethanidine, clonidine,
debrisoquine, deserpidine, guanethidine, guanoclor, guanoxan,
hydralazine, indoramin, labetalol, methyldopa, prazosin, reserpine, sodium
nitroprusside
2. Beta-blockers
Atenolol, oxprenolol, pindolol, sotalol, timolol
3. Antiarrhythmic drugs
Amiodarone, disopyramide, mexiletine, procainamide, tocainide, quinidine
4. Diuretics
Acetazolamide, amiloride (+ hydrochlorothiazide). chlorothiazide,
chlorthalidone, ethacrynic acid, furosemide (+ potassium chloride),
hydroflumethiazide (+ spironolactone), mannitol, methyclothiazide,
metolazone, polythiazide, triamterene (+ hydrochlorothiazide +
benzthiazide). xipamide
5. Coronary vasodilators
Dipyridamole, isosorbide, lidoflazine, nifedipine, pentaerythritol
6. Peripheral vasodilators
Cyclandelate, oxpentifylline, phentolamine, phenoxybenzamine,
thymoxamine
7. Vasoconstrictor and migraine therapy
Clonidine, ergotamine tartrate, isometheptene mucate, methysergide,
pizotifen
8. Anticoagulant and antithrombotic
Nicoumalone
9. Hemostatics
Aminocaproic acid

Drugs aeting on the eentral nervous system

1. Antidepressants
Amitriptyline, clomipramine, desipramine, dothiepin, doxepin, imipramine,
iprindole, iproniazid, isocarboxazid, lithium carbonate, maprotiline,
methylphenidate, mianserin, nomifensine, nortriptyline, phenelzine,
protriptyline, tofenacin, tranylcypromine (+ trifluoperazine), trazodone,
trimipramine, viloxazine
2. Analgesics
Acetylsalicylic acid, benorylate, buprenorphine, codeine,
dextropropoxyphene
(+ aspirin, + caffeinel. dihydrocodeine, diflunisal, methadone,
paracetamol, pethidine, pentazocine
3. Sedatives and tranquillizers
Atropine, chlormezanone, clobazam, diazepam, fluspirilene, lorazepam,
oxazepam, potassium clorazepate, prochlorperazine, promethazine
4. Hypnotics
Flurazepam, lormetazepam, pentobarbitone, temazepam, triazolam
5. Anticonvulsants
Carbamazepine, ethotoin, ethosuximide, methoin, methylphenobarbitone,
phenobarbitone, phenytoin, primidone, sulthiame
6. Stimulants
Dexamphetamine
7. Rigidity and tremor controllers
Amantadine, benzhexol, biperiden, bromocriptine, carbidopa, levodopa,
methixene, orphenadrine, procyclidine
8. Antiemetics
Cyclizine, meclozine (+ pyridoxine)

Muscle relaxants

Baclofen, chlormezanone, dantrolene, methocarbamol, orphenadrine (+


paracetamol)

Non-steroidal anti-inflammatory agents

Dextropropoxyphene (+ aspirin). diclofenac, fenbufen, fenoprofen,


flurbiprofen, hydroxychloroquine, ibuprofen, indomethacin, naproxen,
phenylbutazone, sulindac

Hormones

1. Corticosteroids and related drugs


Cortisone acetate, dexamethasone, fludrocortisone, hydrocortisone,
methylprednisolone, prednisolone, triamcinolone
2. Drugs in diabetes
Fenfluramine, glipizide, insulin, tolazamide

3. Gonadal hormone and related compounds


Cyproterone, danazol, medroxyprogesterone, oestradiol valerate (+
levonorgestrel)
4. Hyper- and hypoglycaemics
Chlorpropamide
5. Trophic hormones
Clomiphene citrate

Drugs used in the treatment of infections and infestations

1. Systemic antibiotics
Amikacin, ampicillin (+ cloxacillin, + flucloxacillin), cephalexin, colistin,
erythromycin, gentamicin, minocycline, netilmicin, pivampicillin, polymixin
B, spectinomycin, streptomycin, tobramycin
2. Topical antibiotics
Neomycin (+ zinc bacitracin)
3. Sulphonamides
Sulphamethoxypyridazine, sulphapyridine
4. Antituberculous drugs
Capreomycin sulphate, cycloserine, isoniazid (+ para-aminosalicylatel.
rifampicin
5. Anthelmintic
Thiabendazole
6. Antimalarial
Chloroquine
7. Antifungal
Amphotericin

Drugs aeting on the respiratory system

1. Expectorants
Isoaminile, triprolidine + pseudo-ephedrine + codeine
2. Bronchospasm relaxant
Aminophylline + ephedrine, deptropine + isoprenaline, isoetharine,
ketotifen
3. Respiratory stimulant
Doxapram

Anti-allergie drugs

Chlorpheniramine, clemastine, cyproheptadine, dimethothiazine,


diphenylpyraline, ephedrine, mepyramine, pethidine, promethazine +
atropine, triprolidine, trifluoperazine

Oral contraeeptives

(Combined and progestogen only)

Carcino-ehemotherapeutic drugs

Cisplatin, hydroxyurea, vinblastine

Gout treatment

Allopurinol, probenecid

Anaestheties, premedication
Lignocaine

Miseellaneous

1. Prostaglandins
Dinoprost, dinoprostone
2. X-ray contrast media
Meglumine and/or iothalamate, diatrizoate
3. Thyrotrophin-releasing hormone
TRH-Roche
4. Glaucoma drug
Dichlorphenamide
5. Treatment of obesity
Fenfluramine, mazindol
6. Treatment of hypoparathyroidism
Dihydrotachysterol (AT 10)
7. Antidiarrhoeal
Diphenoxylate (+ atropine)
8. Treatment of peptic ulceration
Cimetidine, propantheline
9. Treatment of ulcerative colitis
Sulphasalazine

397

Drugs and vertigo

activation is not confined to the liver (Crann and


Schacht 1996).
The reversible ototoxicity of high -dose salicylate
(aspirin) is characterised by bilateral hearing loss
(elevation of pure tone threshold), tinnitus, and vertigo. This toxicity is related to plasma salicylate
levels, which suggests that the underlying cause is an
alteration in the biochemical or enzymatic activity
of the hair cells and cochlear neurons (Silverstein et
al. 1967). No morphological changes could be
demonstrated in the inner ear of squirrel monkeys
after acute intoxication (Myers and Bernstein 1965).
The effects of salicylates on hearing may arise from
the partitioning of the salicylate molecule in the
membrane of the outer hair cells which consequently
inhibits outer hair cell motility (Tunstall et al. 1995)
as well as alters mechanico-electrical transductions
(Aoyagi et al. 1996). Fortunately, all decrements in
cochlear and vestibular function are reversible within days after ceasing salicylate ingestion.
The laap diuretics ethacrynic acid and furosemide
are ototoxic (Schwartz et al. 1970; Mathog et al. 1970)
and cause damage to the stria vascularis (Quick and
Duvall 1970; Jonsson and Hawkins 1972) and corresponding areas of the vestibular labyrinth. The
access of furosemide to its site of ototoxic action
depends on the quantity of unbound furosemide in
the serum (Whitworth et al. 1993). These side effects
are commonly reversible. However, when aminoglycosides are combined with loop-inhibiting diuretics,
the damage may be permanent.
The alkaloid quinine (p. 356) - mainly used for
malaria, cardiac arrhythmia, and nocturnal muscle
cramps - causes reversible or irreversible ototoxic
side effects due to vasospasms or a direct blockade
of the ion channels (Roden 1996; Tracy and Webster
1996; Strupp et al. 1998).
Alkylating chematherapeutic (anticancer) agents
can cause irreversible cochlear and vestibular loss
due to toxie degeneration of the hair cells
(Schuknecht 1974). Attempts to prevent the ototoxicity of certain drugs by simultaneously applying
others have been unsuccessful (Ballantyne 1984). In
vitro and animal studies have provided evidence
that sodium thiosulfate is a potential chemoprotectant (Neuwelt et al. 1996). It has been shown in the
chinchilla vestibular sensory epithelium that carboplatin - a second-generation platinum drug used in
the treatment of cancer - leads to quite rapid
vestibulotoxic effects as a result of significant disruptions of the mitochondria in hair cells and their
afferent terminals (Ding et al. 1997). The degree of
hair cell damage appears to be dependent on the
peak level of carboplatin rather than on the total
dose (Takeno et al. 1994).

Cerebellar intoxication
The antiepileptic diphenylhydantoin is the prototype
of a cerebellotoxic drug. Other toxic compounds
such as alcohol or mercury can affect both the
peripheral vestibular labyrinth and central cerebellar structures.
Cerebellar degeneration has often been reported
to occur in epileptic patients with normal or toxic
serum levels while on long-term phenytain treatment (Ghatak et al. 1976; McLain et al. 1980). Clinical
cerebellar syndrome (Nozue et al. 1973) and cerebellar atrophy following single bouts of phenytoin
intoxication (Lindvall and Nilsson 1984) demonstrate
that this effect is not due to hypoxie events secondary to severe seizures (Spielmeyer 1920). Acute
intoxication manifests as a prolonged, severe disturbance of equilibrium, ataxia, and postural imbalance
with apredominant 3 Hz fore-aft body sway (Fig.
16.5), which resembles that of late alcohol anterior
lobe atrophy (Dichgans et al. 1976). Concurrent ocular motor abnormalities resemble those of experimental ablation of the flocculus in primates (Fig.
28.1; p. 398). The pathomechanism of phenytoin toxicity may involve 3'-5'cycloguanosine monophosphate (Mao et al. 1975). Phenytoin causes selective
damage of the cerebellar Purkinje cells (Salcman et
al. 1978; Breiden-Arendts and Gullotta 1981) and
irreversible cerebellar degeneration (Kuruvilla and
Bharucha 1997). Phenytoin also induces apoptotic
cell death of cultured rat cerebellar granule neurons
(Yan et al. 1995), inhibits expression of microtubuleassociated protein 2, and influences cell viability and
neurite growth (Kempermann and Volk 1995).
Marked degeneration of the Purkinje celllayer in
chronic intoxication due to long-term phenytoin
therapy appears to be largely compensated for functionally, since obvious cerebellar signs are mild.
Intoxication with the antiepileptic drug carbamazepine causes a reversible cerebellar syndrome
(Warot et al. 1967; Umeda and Sakata 1977) that is
clinically indistinguishable from phenytoin intoxication. The most common adverse effect of antiepileptic barbiturates is drowsiness, but dizziness and
ataxia (Livingstone and Petersen 1956) as well as eye
movement abnormalities (Tables 28.2 and 28.3) have
been reported. The vestibulo-ocular reflex and pursuit are both transiently sensitive to barbiturates,
whereas their action on saccades and fixation suppression of the vestibulo-ocular reflex are more prolonged (Dayal et al. 1987). The site of action differs
from that of phenytoin and carbamazepine;
it involves activation of post-synaptic GABA A

398

Vertigo

10'

hOrlz

ENG

20'

Table 28.2. Toxic ocular motor disturbances

JIIIIOI/ll

, , , ,

20.2.1979 (. .,.... level 130

30'

40'

gaze rlght

~----.....,....-----

20'

smooth PIIsuit (0.2 Hz . 60' amplitude)

Ocular motor sign

Toxic agent

Gaze-evoked nystagmus or
saccadic pursuit

Phenytoin, carbamazepine,
barbiturates, alcohol,
benzodiazepines, methadone,
chloral hydrate, marijuana

Siow saccades

Barbiturates, alcohol,
benzodiazepines, fentanyl,
vestibular sedatives
(dimenhydrinate, scopolamine,
cholinergic and anticholinergic
drugs)

Impaired VOR

Vestibular sedatives
(dimenhydrinate, scopolamine),
barbiturates, alcohol,
benzodiazepines

Exaggerated VOR

Industrial solvents (xylene, styrene,


trichloroethylene,
methyl chloroform)

Internuclear ophthalmoplegia

Phenytoin, barbiturates, tricyclic


antidepressants, bromides, hepatic
coma

Externalophthalmoplegia

Phenytoin, barbiturates, tricyclic


antidepressants (imipramine,
doxepin, amitriptyline)

Vertical gaze palsy

Barbiturates/primidone

Skew deviation

Lithium, hepatic coma

Downbeat nystagmus/vertigo

Phenytoin, barbiturates, lithium,


alcohol, magnesium depletion,
vitamin B12 deficiency

Upbeat nystagmus/vertigo

Antiepileptics, tobacco (nicotine)

Periodic alternating nystagmus

Phenytoin

OKN (SO'/s)

~~
VOR fix suPP- (0.1 HZ.60'/s)

5.4.1879 (20 lI/molJl)


20'

smooth pursuit

OKN (SO'/s)

~~
VOR fix supp.

55

Fig.28.1. Original recordings of horizontal eye movements in


acute phenytoin intoxication in a 17-year-old female patient
(serum level: 130 ].lmol/I). Lateral gaze of 10-40 shows increasing gaze-evoked nystagmus. Triangular pursuit (60 amplitude at
0.2 Hz) is saccadic, OKN (stimulus velocity: 60/5) is reduced, and
fixation suppression of the VOR (sinusoidal body oscillations
60/5 at 0.1 Hz) is impaired (top). When the serum level of phenytoin was down to 20 ].lmol/I 6 weeks later, eye movements
appeared almost normal (bottom).

receptors, thereby increasing chloride conductance


and depression of post-synaptic glutamate-induced
excitation (Barker and Ransom 1978).
Cerebellar ataxia has been reported in tumour
patients to be a paraneoplastic syndrome (Auth and
Chodoff 1957; Brain and Henson 1958) as well as a
reversible side effect of chemotherapeutic agents
such as 5-fluorouracil, a chemotherapeutic agent for
palliative treatment of advanced cancer (Riehl and
Brown 1964; Boileau et al. 1971).
Industrial solvents are known to produce a psychogenic syndrome, dizziness, and postural imbalance in humans, as weIl as positional nystagmus and
exaggerated rotatory nystagmus responses in rabbits
(Aschan et al., 1977). This has been described after
exposure to xylene, styrene, trichloroethylene, and
methylchloroform (Larsby et al. 1978 a,b; Ballantyne

Acquired pendular nystagmus

Toluene

Central positional nystagmus

Barbiturates, mercury compounds,


industrial solvents (xylene, styrene,
trichloroethylene,
methylchloroform)

Labyrinthine positional nystagmus

Alcohol, glycerol, "heavy" water

Opsoclonus, ocular flutter

Amitriptyline, haloperidol, lithium,


thallium, chlordecone, DDT,
toluene, cocaine

and Ajodhia 1984; Larsby et al. 1986; Ledin et al.


1989). The induced positional nystagmus is not
caused by peripher al labyrinthine action (see
Buoyancy Hypothesis; p. 285), because the direction
of beating is opposite to that of positional alcohol
nystagmus (Aschan et al. 1977); however, the specific
weight of xylene is also lower than that of
endolymph. Positional nystagmus, exaggerated rotatory nystagmus, and decreased ability to suppress
the VOR visually suggest that these agents block the
inhibitory cerebellar Purkinje efferences on the VOR
(Tham et al. 1979; Haglid et al. 1981; Tham et al.

Drugs and vertigo

399

Table 28.3. Oeular motor abnormalities eaused by toxie agents


Drugs and toxie agents

Oeulomotor abnormalities

Phenytoin

Gaze-evoked nystagmus, saeeadie


pursuit, external ophthalmoplegia,
internuclear ophthalmoplegia;
Rare: periodie alternating nystagmus,
down beat nystagmus/vertigo

Carbamazepine

Gaze-evoked nystagmus, saeeadie


pursuit, impairment of optokinetie
nystagmus, external ophthalmoplegia,
downbeat nystagmus/vertigo, oeulogyrie
crises

Barbiturates/primidone

Gaze-evoked nystagmus, saeeadie


pursuit, impairment of optokinetie
nystagmus, slow saeeades, vertieal gaze
palsy, external ophthalmoplegia,
impairment ofVOR, internuelear
ophthalmoplegia, eentral positional
nystagmus, impaired vergenee,
decreased aecommodative
eonvergenee/aeeommodation ratio

Alcohol

Labyrinthine positional vertigo, saeeadie


pursuit, slow saeeades, down beat
nystagmus/vertigo

Benzodiazepines

Saeeadie pursuit, slow saeeades,


impairment ofVOR, impairment of gazeholding

Trieyelie antidepressants

Internuclear ophthalmoplegia, external


ophthalmoplegia,opsoclonus

Trieyclie antidepressants +
L-tryptophan

Oeular oscillations and myoclonus

Bromides

Internuclear ophthalmoplegia

Vestibular sedatives

Impairment ofVOR, slow saeeades

Lithium

Alternating skew deviation, opsoelonus,


down beat nystagmus/vertigo

Thallium

Opsoclonus

Phenothiazines

Internuclear ophthalmoplegia,
oeulogyrie crises

Methadone

Saeeadie pursuit, saeeadie hypometria

Haloperidol

Opsoclonus, oeulogyrie crisis

Marijuana

Gaze-evoked nystagmus, saeeadie


pursuit

Amphetamine

Inereased aeeommodative convergenee/


aecommodation ratio

Mereury

Spontaneous and/or positional


nystagmus, impairment of optokinetie
nystagmus

Chemotherapeutie anticaneer agents Vestibular/hearing 1055


Loop diureties (ethacrynie acid,
furosemide)

Transient vestibular/hearing 1055,


impaired VOR

Aminoglyeoside antibioties
(gentamicin, streptomycin)

Permanent vestibular and hearing 1055

Aeetylsalieylie acid

Transient vestibulocoehlear impairment

Industrial solvents (xylene,


triehloroethylene, toluene)

Central positional nystagmus,


exaggerated VOR, impaired fixation
suppression ofVOR, opsoclonus,
aequired pendular nystagmus

1984; Larsby et al. 1986). Toluene is one of the most


widely abused organic solvents. Evidence has accumulated on its deleterious effect on the auditory and
vestibular systems (Morata et al. 1995). Moreover, the
loss of outer hair cells of the organ of Corti has been
demonstrated in rats treated with toluene (Campo et
al. 1997). Chronic inhalation leads to a progressive
syndrome with cognitive impairment, cerebellar
ataxia, and abnormal eye movements, including
opsoclonus and acquired pendular nystagmus (Maas
et al. 1991). This has also been experimentally tested
in the rat (Nylen et al. 1991).
"Minamata disease" (Hunter-Russel syndrome), a
cerebellar disequilibrium related to ingestion of
seafood contaminated by mercury compounds in
Minamata Bay during the 1950s, was described by
Kurland et al. (1960), Tokuomi et al. (1961), and
Takeuchi et al. (1962). A similar endemie dis aster of
chronic poisoning by organic mercurials recurred in
1965 to inhabitants in the riverside area of the
Aganogawa river. The disease begins with progressive numbness of the distallimbs, perioral parts, and
the tongue. Ataxia, deafness, dizziness, unsteadiness
of gait, speech disturbance, and worsening of vision
follow (Tsubaki et al. 1969). Ataxia is a major sign,
which is most pronounced for the fore-aft body sway
(Ozawa et al. 1985). Spontaneous and positional nystagmus and impaired horizontal and vertical optokinetic nystagmus have frequently been observed in
these patients (Mizukoshi et al. 1975). A follow-up
study (Mizukoshi et al. 1989) remained inconclusive
in view of the contradietory neuro-otological findings on improvement or deterioration of single signs
and symptoms. Neuro-otological disturbances in
patients with chronie mercurial intoxication were
attributed mainly to retrocochlear and retrovestibular brainstem and cerebellar structures (Mizukoshi
et al. 1975). Damage of the peripherallabyrinthine
epithelium, adjacent nerve endings, and the secretory epithelium was also observed in animals after
mercury chloride intoxication (Anniko and Sarkady
1978).
Otoneurological signs of acute alcohol intoxication include positional alcohol nystagmus (p. 286),
spinocerebellar and vestibulocerebellar ataxia, and
ocular motor abnormalities (Fig. 28.2; Tables 28.2,
28.3). Alcohol, for example, has a slowing effect on
the latency and velocity of voluntary saccadic eye
movements (Katoh 1988), the gain of the vestibuloocular reflex (Takohashi et al. 1989), and postural
control (Ledin and dkvist 1991). Chronic alcohol
intoxieation and malnutrition cause a classic clinical
syndrome (late cortical cerebellar atrophy or anterior
lobe syndrome) characterised by predominant spinocerebellar ataxia of gait and stance with mild or
absent ataxia of the upper limbs (Diehgans and

400

Vertigo

mgll00ml
20

o
63
123

20.

L~

~~
smooth pU,suit (40 ampltude )

0.25 Hz

o
63

123

63
123

0.5 Hz

V\/'\./" 0MM
V'\/V" ~
\J\JV"- VVV\fV\
VOR (0.25 Hz. 6O/s) eyes open

fix suPP

----~...

5s

Fig. 28.2. Original recordings of horizontal eye movements


during acute alcohol intoxication of increasing severity in a 32year-old male patient (from 0 to 63 to 123 mg/1 00 ml). Gazeevoked nystagmus occurs at 123 mg/100 ml serum level (top).
Saccadic pursuit can be observed at serum levels of 63
mg/100 ml (middle) as weil as impaired fixation suppression of
the VOR at sinusoidal body oscillations of 60 0 /s at 0.25 Hz
(bottom). (From Esser and Brandt 1983.)

Diener 1986). The increased body sway is an oscillatorlike 3-Hz sway, predominantly in the fore-aft direction (Silfverskild 1969; Diehgans et al. 1976, Fig.
16.5). Follow-up studies in patients with late atrophy
of the anterior lobe show that there is significant
improvement of postural balance in those who
abstain from such drugs (Vietor et al. 1959; Diener et
al.1984).

Drugs and eye movements


Pathologieal eye movements such as gaze-evoked
nystagmus or saccadie pursuit are typieal but not
sensitive signs of intoxication. They are unspecific
and do not allow identification of the provo king
agent. Among the variety of drug-induced patterns
of pathologieal eye movements, (Tables 28.2 and

28.3) special emphasis should be placed on hydantoins, carbamazepine, barbiturates, benzodiazepines,


amitriptyline, and alcohol (Esser and Brandt 1983;
Brandt and Bchele 1983; Leigh and Zee 1991;
Straube and Brandt 1998). These substances may
have analogous and/or distinct sites of action within
the labyrinths, brainstem, and cerebellum (Table
28.4).
Most of the oculomotor disturbances are possibly
due to a pharmacologically induced transient dysfunction of the vestibulocerebellar-fl.occulus loop,
whieh causes disinhibition of the oculomotor
system, impairment of OKN and fixation suppression of an exaggerated VOR (Esser and Brandt 1983).
Bilateral fl.occulectomy in monkeys causes a pattern
of eye movement signs whieh is also seen during
pharmacologieal intoxication in humans (Figs. 28.1
and 28.2). It consists of saccadie pursuit, gazeevoked nystagmus, impaired OKN, normal or exaggerated VOR, impaired fixation suppression ofVOR,
and downbeat nystagmus (Takemori and Cohen
1974; Takemori and Suzuki 1977; Zee et al. 1978,
1981).
Patients with the clinical syndrome of intoxieation present with other cerebellar dysfunctions such
as dysarthria, ataxia, and postural imbalance. Toxie
ocular oscillations (Tables 28.2 and 28.3) as weIl as
ocular deviations or ophthalmoplegia may cause
diplopia, blurred vision, and apparent movement of
the visual scene (oscillopsia) and thereby contribute
to distressing disequilibrium of stance and gait
(Visual vertigo; p. 430).
Vestibular sedatives such as dimenhydrinate or
scopolamine (TTS-scopolamine) not only prevent
motion siekness by selective suppression ofvestibular nuc1ei neurons (decreased nystagmus response to
rotatory and calorie stimuli; Fig. 28.3) but also
impair optokinetie nystagmus, reduce maximal
velo city of saccades, cause considerable central sedation, and prolong central vestibular compensation
(Brandt et al. 1974; Pyykk et al. 1984,1985).
A single agent can cause several distinct or combinations of oculomotor disturbances. For example,
alcohol causes labyrinthine dysfunction with positional nystagmus (Aschan 1958), a cerebellar eye
movement pattern of gaze-evoked nystagmus, saccadic pursuit, impaired fixation suppression of the
VOR (Fig. 28.2; Esser and Brandt 1983), reduced
effectiveness in the visual feedback of retinal error
information during head movements (Barnes et al.
1985), transient downbeat nystagmus in cases of
acute intoxieation (Rosenberg 1987), permanent
downbeat nystagmus with chronie cerebellar degeneration (Zasorin et al. 1984), and spinocerebellar
ataxia in the anterior lobe syndrome (Dichgans and
Diener 1986).

401

Drugs and vertigo


Table 28.4. Toxic agents inducing ocular motor or vestibular disturbanees: predominant site of action

Brainstem/cerebellum
Barbiturates/primidone
Benzodiazepines
Bromides
Carbamazepine

Labyrinth

Chlordecone

Aminoglycoside antibiotics

Chloral hydrate

(gentamicin, streptomycin)

Phenytoin

Loop diuretics (ethacrynic acid,

Haloperidol

furosemide)

Tricyclic antidepressants

Salicylates

Lithium

Chemotherapeutic anticancer agents

Marijuana
Methadone
Phenothiazines
Thallium
Vestibular sedatives

20

90%

rolalion olop

horiz.
ENG

6:]-'"--~-W
"",",,,,,,,,,,,,,,
J

1
150

.r.N',,","'\.!"'<'

100mg
DIMENHYDRINA TE

ci' 2 I
R

I~:j---~-I
15

-----!
l..__----------------"-.,--.--0,6mg
SCOPOlAMINE

3 sec
----...

Fig.28.3. Original recordings of post-rotatory nystagmus, following a rapid stop (15) from a constant velocity body rotation at 90/5,
in two normal subjects with and without administration of 100 mg dimenhydrinate (top) or 0,6 mg scopolamine (bottom). Both drugs
largely suppress the post-rotatory nystagmus, an effect that is used to treat vertigo and nausea in acute clinical vertigo syndromes or
to prevent motion sickness, (From Brandt and Bchele 1983.)

On the other hand, different agents can cause one


particular syndrome. Downbeat nystagmus (p. 199),
for example, is caused by alcohol (Zasorin and Baloh
1984; Rosenberg 1987), lithium (Halmagyi et al.
1989; Corbett et al. 1989), phenytoin (Alpert 1978),
carbamazepine (Wheeler et al. 1982), magnesium
depletion (Saul and Selhorst 1981), amiodarone

(Arbusow et al. 1998), or vitamin Bl2 deficiency


(Mayfrank and Thoden 1986). Upbeat nystagmus
(p.205) can also be elicited by pharmacological
intoxications due to antiepileptics or the excitatory
effects of nicotine in tobacco smoking (Sibony et al.
1987).

402

References
Alpert JN (1978) Downbeat nystagmus due to anticonvulsant toxicity. Ann Neuro14:471-473
Anniko M, Sarkady L (1978) The effects of mercurial poisoning on
the vestibular system. Acta Otolaryngol (Stockh) 85:96-104
Aoyagi M, Yoshida M, Makishima K (1996) Different effects of
noise and salicylate and their interaction on the guinea pig
cochlea. Eur Arch Otorhinolaryngo1253:429-434
Arbusow V, Strupp M, Brandt Th (1998) Amiodarone-induced
severe, prolonged head-positional vertigo and vomiting.
Neurology 51:917
Aschan G (1958) Different types of alcohol nystagmus. Acta
Otolaryngol (Stockh) SupplI40:69-78
Aschan G, Bunnfors I, Hyden D, Larsby B, dkvist LM, Tham R
(1977) Xylene exposure. Electronystagmographic and gas chromatographic studies in rabbits. Acta Otolaryngol (Stockh)
84:370-376
Auth TL, Chodoff P (1957) Transient cerebellar syndrome from
extracerebral carcinoma. Neurology 7:370
Ballantyne JC (1970) latrogenic deafness. J Laryngol Otol
84:967-1000
Ballantyne J (1984) Ototoxicity. In: Oosterveld WJ (ed)
Otoneurology. Wiley, Chi chester, pp 41-51
Ballantyne J, Ajodhia J (1984) latrogenic dizziness. In: Dix MR,
Hood JD (eds) Vertigo. Wiley, Chichester, pp 217-247
Barker JL, Ransom BR (1978) Pentobarbitone pharmacology of
mammalian central neurones grown in tissue culture. J Physiol
(Lond) 280:355-372
Barnes GR, Crombie JW, Edge A (1985) The effects of ethanol on
visual-vestibular interaction during active and passive head
movements.Aviat Space Environ Med 56:695-701
Black FO, Pesznecker SC (1993) Vestibular ototoxicity. Clinical
considerations. Otolaryngol Clin North Am 26:713-736
Boileau G, Piro AJ, Lahiri SR, Hall TC (1971) Cerebellar ataxia during 5-fluorouracil (NSC-19893) therapy. Cancer Chemother Rep
55:595-598
Brain WR, Henson RA (1958) Neurological syndromes associated
with carcinoma. Lancet II:971
Brandt Th, Bchele W (1983) Augenbewegungsstrungen. Klinik
und Elektronystagmographie. Fischer, Stuttgart
Brandt Th, Dichgans J, Wagner W (1974) Drug effectiveness on
experimental optokinetic and vestibular motion sickness.
Aerospace Med 45:1291-1297
Breiden-Arendts D, Gullotta F (1981) Diphenylhydantoin,
Epilepsie, Kleinhirnatrophie - Histologische und elektronenmikroskopische Untersuchungen. Fortschr Neurol Psychiatr
49:406-414
Campo P, Lataye R, Cossec B, Placidi V (1997) Toluene-induced
hearing loss: a mid-frequency location of the cochlear lesions.
Neurotoxicol TeratolI9:129-140
Corbett JJ, Jacobson DM, Thompson HS, Hart MN, Albert DW
(1989) Downbeating nystagmus and other ocular motor defects
caused by lithium toxicity. Neurology 39:481-487
Crann SA, Schacht J (1996) Activation of aminoglycoside antibiotics to cytotoxins.Audiol Neurootoll:80-85
Cruickshank JM (1981) Beta-blockers, bradycardia and adverse
effects. Acta Therapeut 7:309-321
Currier WD (1970) Dizziness related to hypoglycemia: The role of
adrenal steroids and nutrition. Laryngoscope 80: 18-35
Dayal VS, Mai M, Tomlinson RD, Farkashidy J (1987) Effects of
barbiturate on the vestibular and oculomotor systems: a
sequential study. In: Graham MD, Kemink JL (eds) The vestibular system: neurophysiologic and clinical research. Raven Press,
New York, pp 169-175
Dichgans J, Mauritz KH, Allum JHJ, Brandt Th (1976) Postural
sway in normals and atactic patients: analysis of the stabilizing
and destabilizing effects of vision. Agressologie 17: 15-24

Vertigo
Dichgans J, Diener HC (1986) Different forms of postural ataxia in
patients with cerebellar diseases. In: Bles W, Brandt Th (eds)
Disorders of posture and gait. Elsevier,Amsterdam, pp 207-215
Diener HC, Dichgans J, Bacher M, Guschlbauer B (1984)
Improvement of ataxia in late cortical cerebellar atrophy
through alcohol abstinence. J Neuro1231:258-262
Ding D, Wang J, Salvi RJ (1997) Early damage in the chinchilla
vestibular sensory epithelium from carboplatin. Audiol
Neurooto12:155-167
Esser J, Brandt Th (1983) Pharmakologisch verursachte
Augenbewegungsstrungen.
Differentialdiagnose
und
Wirkungsmechanismen. Fortschr Neurol Psychiatr 51 :41-56
Ghatak NR, Sontoso RA, McKinney WM (1976) Cerebellar degeneration following long-term phenytoin therapy. Neurology
26:818-820
Haglid KG, Briving C, Hansson HA, Rosengren L, Kjellstand P,
Stabron D, Swedrin U, Wronski A (1981) Trichloroethylene: long
lasting changes in the brain after rehabilitation.
Neurotoxicology 2:659-673
Hain TC, Zee DS (1991) Abolition of optokinetic afternystagmus
by aminogycoside ototoxicity. Ann Otol Rhinol Laryngol
100:580-583
Halmagyi GM, Lesell I, Curthoys IS, Lessel S, Hoyt WF (1989)
Lithium-induced downbeat nystagmus. Am J Ophthalmol
107:664-670
Harrill JA (1951) Headache and vertigo associated with hypoglycemic tendency. Laryngoscope 61: 138-145
Herdman SJ, Sandusky AL, Hain TC, Zee DS, Tusa RJ (1994)
Charateristics of postural stability in patients with aminoglycoside toxicity. J Vestib Res 4:71-80
Hinshaw HC, Feldman WH (1945) Streptomycin in treatment of
clinical tuberculosis: A preliminary report. Proc Staff Meet
Mayo Clin 20:313
Jonsson LG, Hawkins JE jr (1972) Strial atrophy in clinical and
experimental deafness. Laryngoscope 83:1105-1125
Jonsson LG, Wright CG, Preston RE, Henry PJ (1980)
Streptomycin-induced defects of otoconial membrane. Acta
Otolaryngol (Stockh) 89:401-406
Katoh Z (1988) Slowing effects of alcohol on voluntary eye movements. Aviat Space Environ Med 59:606 -61 0
Kempermann G, Volk B (1995) Phenytoin inhibits expression of
microtubule-associated protein 2 and influences cell-viability
and neurite growth of cultured cerebellar granule cells. Brain
Res 687:194-198
Kurland L, Faro SN, Siedler H (1960) Minamata disease: the outbreak of a neurologic disorder in Minamata, Japan, and its relationship to the ingestion of seafood contaminated by mercuric
compounds. World Neuroll:370-391
Kuruvilla T, Bharucha NE (1997) Cerebellar atrophy after acute
phenytoin intoxication. Epilepsia 38:500-502
Larsby B, Tham R, dkvist LM, Hyden D, Bunnfors I, Aschan G
(1978a) Exposure of rabbits to styrene. Electronystagmographic findings correlated to the blood and CSF styrene
levels. Scand J Work Environ Health 4:60-65
Larsby B, Tham R, dkvist LM, Norlander B, Aschan G, Rubin A
(1978b) Exposure of rabbits to methylchloroform. Vestibular
disturbances correlated to blood and CSF levels. Int Arch Occup
Environ Health 41:7-15
Larsby B, Tham R, Eriksson B, Hyden D, dkvist L, Liedgren C,
Bunnors I (1986) Effects of trichloroethylene on the human
vestibulo-oculomotor system. Acta Otolaryngol (Stockh)
101:193-199
Ledin T, dkvist LM, Mller C (1989) Posturography findings in
workers exposed to industrial solvents. Acta Otolaryngol
(Stockh) 107:357-361
Ledin T, dkvist LM (1991) Effect of alcohol measured by dynamic
posturography. Acta Otolaryngol (Stockh) Supp1481:576-581
Leigh RJ, Zee DS (1991) The neurology of eye movements, 2nd
edn. Davis, Philadelphia
Lindeman H (1969) Regional differences in sensitivity of the

Drugs and vertigo


vestibular sensory epithelia to ototoxic antibiotics. Acta
Otolaryngol (Stockh) 67:177-189
Lindvall 0, Nilsson B (1984) Cerebellar atrophy following phenytoin intoxication. Ann NeuroI16:258-260
Livingstone S, Petersen D (1956) Primidone (mysoline) in the
treatment of epilepsy. Results of treatment of 486 patients and
review on the literature. N Engl J Med 254:327-329
Maas EF,Ashe J, Spiegel P, Zee DS, Leigh RJ (1991) Acquired pendular nystagmus in toluene addiction. Neurology 41:282-285
Mao CC, Guidotti A, Costa E (1975) Inhibition by diazepam ofthe
tremor and the increase of cerebellar cGMP content elieited by
harmaline. Brain Res 83:516-519
Mathog RH, Thomas WG, Hudson WR (1970) Ototoxicity of new
and potent diuretics. Arch Otolaryngol 92:7-13
Mayfrank L, Thoden U (1986) Downbeat nystagmus indicates
cerebellar or brain-stem lesions in vitamin Bi2 -deficiency. J
NeuroI233:145-148
McLain LW Jr,Martin JT,Allen JH (1980) Cerebellar degeneration
due to chronic phenytoin therapy. Ann NeuroI7:18-23
Mizukoshi K, Nagaba M, Ohno Y, Ishikawa K, Aoyagi M, Watanabe
Y, Kato I, Ino H (1975) Neurootological studies upon intoxication by organic mercury compounds. ORL 37:74-87
Mizukoshi K, Watanabe Y, Kobayashi H, Nakano Y, Koide C,
Inomata S, Saitoh H (1989) Neurotological follow-up studies
upon Minamata disease. Acta Otolaryngol (Stockh) Suppl
468:353-357
Morata TC, Nylen P, Johnson AC, Dunn DE (1995) Auditory and
vestibular functions after single or combined exposure to
toluene: a review. Arch Toxicol 69:431-443
Murofushi T, Halmagyi GM, Yavor RA (1997) Intratympanic gentamicin in Meniere's disease: results of therapy. Am J Otol
18:52-57
Myers EN, Bernstein JM (1965) Salicylate ototoxicity. A clinical
and experimental study.Arch LaryngoI82:483-493
Neuwelt EA, Brummett RE, Remsen LG, Kroll RA, Pagel MA,
McCormick CI, Guitjens S, Muldoon LL (1996) In vitro and
animal studies of sodium thiosulfate as a potential chemoprotectant against carboplatin-induced ototoxicity. Cancer Res
56:706-709
Nilges TC, Northern JL (1971) latrogenic ototoxic hearing loss.
Ann Surg 173:281-289
Nozue M, Mizuno M, Kaga K (1973) Neurootological findings in
diphenylhydantoin intoxication. Ann OtoI82:389-394
Nylen P, Larsby B, Johnson A-C, Eriksson B, Hglund G, Tham R
(1991) Vestibular-oculomotor, opto-oculomotor and visual
function in the rat after long-term inhalation exposure to
toluene.Acta Otolaryngol (Stockh) 111:36-43
Ozawa H, Ishikawa S, Mukuno K (1985) Balance study of methyl
mercury poisoning. In: Igarashi M, Black 0 (eds) Vestibular and
visual control on posture and locomotor equilibrium. Karger,
Basel, pp 302-308
Pyykk I, Schalen L,Jntti V, Magnusson M (1984) A reduction of
vestibulo-visual integration during transdermally administered
scopolamine and dimenhydrinate. Acta Otolaryngol (Stockh)
SuppI406:167-173
Pyykk I, Schalen L, Matsuoka I (1985) Transdermally administered scopolamine vs. dimenhydrinate. 11. Effect on different
types of nystagmus. Acta Otolaryngol (Stockh) 99:597-604
Quick CA, Duvall AJ (1970) Early changes in the cochlear duct
from ethacrynic acid: an electron microscopic evaluation.
Laryngoscope 80:954-965
Ramsden RT,Ackrill P (1982) Bobbing oscillopsia from gentamicin toxieity. Br J Audiol16: 147-150
Rascol 0, Hain TC, Brefel C, Benazet M, Clanet M, Montastruc J-L
(1995) Antivertigo medications and drug-induced vertigo. A
pharmacological review. Drugs 50:777-791
Riehl JL, Brown WJ (1964) Acute cerebellar syndrome secondary
to 5-fluorouracil therapy. Neurology 14:961-967
Roden DM (1996) Antiarrhythmic drugs. In: Hardman JG,

403
Limbird LE (eds) Goodman & Gilman's The pharmacological
basis of therapeutics. McGraw-Hili, New York, pp 839-874
Rosenberg ML (1987) Reversible downbeat nystagmus secondary
to excessive alcohol intake. J Clin Neuroophthalmol 7:23-25
Salcman M, Defendini R, Correll J, Gilman S (1978)
Neuropathological changes in cerebellar biopsies of epileptic
patients.Ann NeuroI3:10-19
Saul RF, Selhorst JB (1981) Downbeat nystagmus with magnesium
depletion.Arch NeuroI38:650-652
Schuknecht HF (1974) Pathology of the ear. Harvard University
Press, Cambridge, Mass
Schwartz GH, David DS, Riggio RR, Stenzel KH, Rubin AL (1970)
Ototoxicity induced by furosemide. N Engl J Med
282:1413-1414
Sibony PA, Evinger C, Manning KE (1987) Tobacco-induced
primary-position upbeat nystagmus.Ann NeuroI21:53-58
Silfverskild BP (1969) Romberg's test in the cerebellar syndrome
occurring in chronic alcoholism. Acta Neurol Scand 45:292-302
Silverstein H, Bernstein H, Davies D (1967) A biochemical and
electrophysiological study. Ann Otol Rhinol Laryngol
76:118-128
Spielmeyer W (1920) ber einige Beziehungen zwischen
Ganglienzellvernderungen und glisen Erscheinungen, besonders im Kleinhirn. Z Ges Neurol Psychiatr 54:1-38
Straube A, Brandt Th (1998) Pharmakologisch induzierte
Augenbewegungsstrungen und Pharmakotherapie von
Augenbewegungsstrungen. In: Huber A, Kmpf D (eds)
Klinische Neuroophthalmologie. Thieme, Stuttgart, pp 609-617
Strupp M, Suckfll M, Schwark Ch, Knig G, Brandt Th (1998)
Akute Chinin-Intoxikation: blind, taub und schwindlig. Mnch
med Wochenschr 140:79-81
Takahashi M, Akiyama I, Tsujita N, Yoshida A (1989) The effect of
alcohol on the vestibulo-ocular reflex and gaze regulation. Arch
OtorhinolaryngoI246:195-199
Takemori S, Cohen B (1974) Loss ofvisual suppression ofvestibular nystagmus after flocculus lesions. Brain Res 72:213-224
Takemori T, Suzuki M (1977) Cerebellar contribution to oculomotor function. ORL 39:209-217
Takeno S, Harrison RV, Mount RJ, Wake M, Harada Y (1994)
Induction of selective inner hair cell damage by carboplatin.
Scanning Microsc 8:97-106
Takeuchi T, Morikawa N, Matsumoto H, Shiraishi Y (1962) A
pathological study of Minamata disease in Japan. Acta
NeuropathoI2:40-57
Tham R, Larsby B, Odkvist LM, Norlander B, Hyden D, Aschan G,
Bertler A (1979) The influence of trichloroethylene and related
drugs on the vestibular system. Acta Pharmacol Toxicol
44:336-342
Tham R, Bunnfors I, Eriksson B, Larsby B, Lindgren S, dkvist LM
(1984) Vestibulo-ocular disturbances in rats exposed to organic
solvents. Acta Pharmacol ToxicoI54:58-63
Tokuomi H, Okajima T, Kanai J, Tsunoda M, Ichivasu Y, Misumi H,
Shimomura K, Tabata M (1961) Minamata disease. World
Neurology 2:536-546
Tracy JW, Webster LT (1996) Drugs used in the chemotherapy of
protozoal infections. Malaria. In: Hardman JG, Limbird LE (eds)
Goodman & Gilman's The pharmacological basis of therapeutics. McGraw-Hili, New York, pp 965-985
Tsubaki T, Shirakawa K, Kanbayashi K, Hirota K (1969) Clinical
features of organic mercury intoxication in the Agano area. Adv
Neurol Sei 13:85-88
Tunstall MJ, Gale JE, Ashmore JF (1995) Action of salicylate on
membrane capacitance of outer hair cells from the guinea-pig
cochlea. J Physiol (Lond) 485:739-752
Umeda Y, Sakata E (1977) Equilibrium disorder in carbamazepine
toxieity.Ann Otol Rhinol LaryngoI86:1-5
Victor M, Adams RD, Mancall EL (1959) A restricted form of cerebellar cortical degeneration occurring in alcoholic patients.
Arch NeuroI1:579-688

404
Warot P, Arnott G, Delahousse J, Petit H (1967) Ataxie aigne apres
ingestion massive de tegretal: evolution favorable. Lilie Med
12:601-604
WerslI J (1995) Ototoxic antibiotics: a review. Acta Otolaryngol
(Stockh) SuppI519:26-29
West BA, Brummett RE, Hirnes DL (1973) Interaction of
kanamycin and ethacrynic acid. Arch OtolaryngoI98:32-37
Wheeler SD, Ramsay RE, Weiss J (1982) Drug-induced downbeat
nystagmus.Ann NeuroI12:227-228
Whitworth C, Morris C, Scott V, Rybak LP (1993) Dose-response
relationships for furosemide ototoxicity in rat. Hear Res
71:202-207

Vertigo
Yan GM, Irwin RP, Lin SYZ, Weller M, Wood KA, Paul SM (1995)
Diphenylhydantoin induces apoptotic cell death of cultured rat
cerebellar granule neurons. J Pharmacol Exp Ther 274:983-990
Zasorin NL, Baloh RW (1984) Downbeat nystagmus with alcoholic
cerebellar degeneration. Arch Neurol 41: 130 1-1302
Zee DS, Yamazaki A, Gcer G (1978) Ocular motor abnormalities
in trained monkeys with floccular lesions. Soc Neurosci Abstr
4:168
Zee DS, Yamazaki A, Butler PH, Gcer G (1981) Effects of ablation
of flocculus and paraflocculus on eye movements in primate. J
NeurophysioI46:878-899

SECTION J
Non-vestibular (sensory)
vertigo syndromes

The term "non-vestibular (sensory) vertigo syndromes" refers to visual and somatosensory vertigo. In a strict sense such a distinction is incorrect,
because convergence of multisensory inputs at various levels of the central vestibular system sometimes makes it impossible for neurons as weil as for
conscious perception to differentiate clearly among
vestibular, somatosensory, and visual stimulation.
The sensation of self-rotation, for example, during
actual body acceleration to the right (by use of a
rotating chair) is indistinguishable from the sensation of circularvection (p. 409), whieh purely optokinetie stimulation to the left (visual input) induces
in the stationary subject. Aseparation of visual and
somatosensory vertigo syndromes is, nevertheless,
useful because of the different causes and mechanisms involved.
Visual vertigo (p. 409), the symptoms of which
may include spatial disorientation, apparent
motion (oscillopsia), postural imbalance, and nausea, is induced by either unusual visual stimulation
(physiological visual vertigo) or ocular motor disorders (pathologieal visual vertigo). The most
important visual (stimulation) vertigo syndrome is
height vertigo (p. 418), a physiological postural
instability occurring when the distance between
eye and visual surround becomes critically large.
Furthermore, spatial disorientation may occur
when subjects are exposed to tilted rooms or distorting mirrors, or simply when they wear glasses
that cause unadapted changes in refraction, in ducing a multisensory perceptual confliet. Moving
visual patterns that cover large parts of the visual
field induce apparent self-motion in the direction
opposite to the pattern motion (circularvection,
linearvection, rollvection; p. 411) or optokinetie

motion (simulator) sickness (p. 417). Distressing


tilt sensations accompanied by nausea may occur if
optokinetic stimulation is combined with active
head movements (pseudo-Coriolis effects, pseudoPurkinje effects; p. 416). Visual vertigo with ocular
motor disorders is characterized by acquired ocular
oscillations (p. 432), infranuclear eye muscle paresis
(p. 431), and decrements in visual performance, such
as errors of refraction, diplopia, visual field defects,
or visual hallucination (see Visual (optokinetie)
epilepsy; p. 237).
The most interesting but controversial somatosensory vertigo syndrome is cervieal vertigo
(p. 441), which cannot be diagnosed with our current tools. Experimental ataxia with ipsilateral
extensor muscle hypotonia and deviation of gait
can be induced in humans by unilateral suboccipital anaesthesia of the neck (p. 442) and can produce
positional or spontaneous nystagmus in animals. It
is not yet known, however, whether cervical vertigo
constitutes a clinical entity caused, for example, by
a tone imbalance of upper cervical root input following whiplash injury. Passive movements of the
cervical column (rotation of the trunk with the
head fixed) induce nystagmus and a compelling
illusion of self-motion as do passive movements of
the [imbs in the proximal joints (arthrokinetic nystagmus and self-motion sensation; p. 446). The lack
of somatosensory input, for example in sensory
polyneuropathy (p. 447), manifests in somatosensory ataxia and unsteadiness of gait, partieularly
under conditions of reduced visual input. Finally,
acoustie stimulation with moving sound sources
can also elicit nystagmus and asensation of selfmotion (p. 447), but this sensory modality does not
seem to be relevant in vertigo.

407

Visual vertigo: visual control of motion


and balance

Vision contributes significantly to spatial orientation, self-motion perception, and postural balance.
Therefore, either unusual visual stimulation or visual
sensory dysfunction may cause distressing vertigo
and disequilibrium, such as height vertigo or "visual
ataxia" associated with the sudden onset of an
extraocular eye muscle paresis.
In 1794 Erasmus Darwin, the grandfather of
Charles Darwin, wrote on visual vertigo in his
Zoonomia or The Laws of Organic Life:
"Many people, when they arrive at fifty or sixty
years of age, are affected with slight vertigo; which is
gene rally but wrongly ascribed to indigestion, but in
reality arises from a beginning defect of their
sight. .. These people do not see objects so distinctly
as formerly, and by exerting their eyes more than
usual they perceive the apparent motions of objects,
and confound them with the real motions of them;
and therefore cannot accurately balance themselves
so as easily to preserve their perpendicularity by
them:'
It is historically remarkable that this description
of visual vertigo and its consequences for postural
balance dates back to a time when vestibular function was not yet understood, nearly a century before
Mach's (1875) Grundlinien der Lehre von den
Bewegungsempfindungen. In those days it was generally accepted that body accelerations were sensed by
the distribution of blood and by skin pressure receptors (Henn and Young 1975). Later the term vertigo
was almost completely identified with the vestibular
system, and even nowadays clinicians tend to differentiate between a true (vestibular) and a pseudo
"non-vestibular" vertigo syndrome. This distinction,
however, is not useful since we know that because of
a visual-vestibular convergence, neither secondorder neurons at the level of the vestibular nuclei
nor conscious perception can distinguish between
labyrinthine and optokinetic stimulation (Dichgans
and Brandt 1978).
Vertigo, defined as a displeasing distortion of static
gravitation al orientation or erroneous perception of
either self-motion or object-motion, may thus be

induced by physiological stimulation or pathological


dysfunction of any of the stabilising sensory
systems: vestibular, visual, and somatosensory
(Brandt and Daroff 1980).
Physiological visual vertigo may be induced by
optokinetic stimulation (circularvection, optokinetic
motion sickness) or by a critical distance between
the subject and the nearest visible stationary contrasts (height vertigo), whereas pathological visual
vertigo is either secondary to involuntary eye movements (oscillopsia) or amismatch between the true
and the expected eye position in the head (extraocular muscle paresis). Visual vertigo is a heterogeneous syndrome with peripheral or central
aetiologies and may occur if patients with balance
dis orders show high visual field dependence
(Bronstein 1995). Forms of visual vertigo are summarised in Table 29.1.

Circularvection and linearvection:


optokinetically induced perception of
self-motion
The sensation of self-motion is a common visual
illusion from which inferences concerning visualvestibular interaction may be drawn. The effect may
be perceived while gazing at moving clouds or
streaming water, or when a train moves on the adjacent track in a railway station (circularvection,
linearvection: see Dichgans and Brandt 1978). This
compelling sensation of body movement can even
affect postural balance. Individual differences exist
in the sensitivity to vection illusion, and these differences are reliable across repeated testing sessions
(Kennedy et al. 1996). Sensitivity to visually induced
vection increases with ageing, suggesting that visual
inputs to the perception of self-motion are enhanced
in the elderly (Paige 1994). A first approach to the
problem was a psychological explanation; for example,

409

410

Vertigo

Table 29.1. Forms of visual vertigo (symptoms may include spatial disorientation, apparent motion, oscillopsia, nausea, postural imbalance, gait disturbance)

I. Physiological (non-pothologicol, induced by unusual visual stimulation)


Stimulus
(a).5tatic
Spatial disorientation (illusions)
Distortion or tilt of threedimensional visual environment
(tilted rooms, distorting mirrors)
Height vertigo
Critical distance between eye and
visual surroundings
(Acrophobia)
Critical distance between eye and
visual surroundings
(b) Moving visual fjelds
Linearvection
Circularvection
(pseudo-Coriolis effect)
Rollvection
(pseudo-Purkinje effect)
Optokinetic motion sickness

Visual motion simulator

Motion ahereffects

Prolonged motion stimulation

Linear motion
Angular motion (yaw)
Angular motion (roll)

Mechanism
Mismatch:
- Intravisual
- Vision-otoliths-somatosensors
Loss of visual stabilisation of
posture
Psychogenic (see p. 460)

Mismatch:
- Vision-otoliths
- Vision-horizontal semicircular canals
- Vision-vertical semicircular
canals-otoliths
- Vision-semidrcular canals
otoliths-somatosensors
Motion adaptation, central ahereffects

11. Pathological (by disease or dysfunction)


(a) Visual performance
Error of refraction (bifocal or cataract glasses)
Diplopia
Visual field defects
Visual hallucination (see epileptic vertigo, p. 233)
(b) Ocular motor disorders
Infranuclear ocular motor disorders (extraocular eye muscle paresis, ocular myasthenia, dysthyroid ocular myopathy, etc.)
Acquired ocular oscillations (downbeat/upbeat nystagmus, spasmus nutans, acquired pendular nystagmus, superior oblique myokymia, etc.)

Helmholtz (1896) mentioned "Urteilstuschung".


One might assurne that these illusions are inferences
based on conscious or unconscious assumption of a
stable environment, so that when the environment
does in fact move, the observer infers that he hirns elf
is moving. This interpretation would be consistent
with the individual's past experience, as the entire
visual surroundings seldom move uniformly under
natural conditions unless the body moves relative to
the earth.
However, as in the case of other illusions, physiological explanations may be more rewarding than
psychological ones. As Purkinje suggested, visual
illusions reveal the truth about the visual system
("Gesichtstuschungen sind Sinneswahrheiten"). An
analysis consistent with his view takes into account
that in the visuaHy induced perception of selfmotion, the term illusion applies only to the case in
which the surroundings move exclusively. It also
recognises that identical visual stimulation occurs
during actual body motion, in which instance the
same physiological mechanism may maintain the
veridical perception of body motion at constant
velo city. The assumption of such a mechanism of
visual-vestibular interaction, common to both illusion and reality, may lead to a better understanding

of sensory control of body movement and spatial


orientation than does the psychological explanation.
Under natural conditions, with the eyes open,
active as weH as passive body motion is adequately
perceived, be it of constant or varying velo city (Fig.
29. IB). While one is riding in a vehicle with the eyes
closed, however, the deficiencies of the vestibular
system become apparent. Vestibular information
ab out motion is then evoked only through acceleration or deceleration, and dies out as the cupulae
within the semicircular canals, or the otoliths, progressively return to their resting position during
constant velo city. Consequently, constant velo city in
the dark cannot be discriminated from rest for any
extended period of time. Owing to this lack of sustained labyrinthine information about constant
velo city, deceleration from constant velo city is misinterpreted as acceleration in the opposite direction
(Fig. 29.1A). These two observations support the
assumption that visual inputs are directly integrated
with vestibular afferents, and are able to compensate
for the inability of the vestibular system to respond
correctly to sustained constant velo city input. This
integration of visual and vestibular information may
be an inherent property of the neural architecture at
birth rather than an ability developed solely as a

411

Visual vertigo: visual control of motion and balance

I
2:1t -~I.L' ___________________~~~_
I

20

I
I

II
I

I
---l ---------=-----;I. ,-----t--------

_____
C

I
-------------------1--

I
~
~

~' -.------------------i----, ~
----- ------

----------~~
--- -- -- ----~--------~k

Fig.29.1. Original recordings of continuous tracking of perceived self-motion velocity and direction during motion of chair and/or
surroundings of a large cylindrical drum (trapezoid velocity profile, top trace). During chair rotation in dark (A) velocity profile roughly
follows mechanical characteristics of cupula-endolymph system, resulting in subject's inability to discriminate constant velocity and
consequently to misinterpret deceleration. With visible surroundings providing adequate optokinetic information, these deficiencies
are largely compensated for (B). Net visual effect is demonstrated in (C) where (with considerable latency) apparent self-rotation is
elicited in a stationary observer by means of the exclusive motion of surroundings in the opposite direction.lf visual surroundings
move with observer as in vehicles (D), motion perception is again erroneous since, as in (A), it relies exclusively on vestibular inputs.
(From Dichgans and Brandt 1978.)

result of learning and experience. Thus, a subject's


illusion of deceleration as acceleration in the dark
(see also Fig. 29.1C) may in fact indicate a functionally
important mechanism of multisensory interaction.
Visually induced vection is essential for adequate
perception of self-motion, especially during transportation in vehic1es. Vection illusions may cause
falls in healthy subjects or raise Bight safety concern
among helicopter pilots while Bying over water and
under dark-light conditions (Ungs 1989). Vection is
an important component of the compelling experience of "presence" that users of virtual reality systems have in a virtual environment (Kennedy et al.
1996). However, exposure to virtual environments
often causes users to experience postural disequilibrium and symptoms of motion sickness (Kennedy
and Stanney 1996).

Psychophysics of circularvection
Unlike vestibular stimuli, which invariably lead to
the sensation of body motion, visual motion stimuli

allow two perceptual interpretations: either selfmotion or object-motion (Fig. 29.2). The subject
watching moving stimuli may perceive hirns elf as
being stationary in space (egocentric motion perception) or may experience the actually moving surroundings as being stable and hirns elf as being
moved (circularvection, linearvection). Uniform
motion filling the entire visual field invariably leads
to a self-motion sensation that is indistinguishable
from an actual body movement. Following the onset
of the stimulus, circularvection begins after a few
seconds of latency and slowly increases in apparent
velo city until saturation; it may outlast the visual
stimulus as an aftereffect. The velo city of apparent
self-rotation in circularvection matches stimulus
speed up to 90-100 /s, but at higher speed circularvection velo city lags behind, resulting in additional egocentric object-motion perception.
In aseries of psychophysical experiments, we
showed that the visual perception of self-motion is
dependent on the density of moving contrasts randomly distributed within the visual field, and on the
total area of the coherent stimulus field (Dichgans
0

Vertigo

412

Perception
of

Stimulus
Velocities

Time Course
(aner Stimulus Onset)

Stimulus Area
and Location
on the Retina

Number of
Moving
Contrasts

Foreground
Background

Object Motion

t
Mixed Selt and
Object Motion

0-3

1 - 10 s

t
SeIt - Motion

Foreg' ound

tfCfu W

> 90-120

from 5-10 s
onwards

/s

Bac"",ound

<

90-120 1,

Fig.29.2. Schematic drawing summarising critical optokinetic stimulus properties wh ich determine wh ether object-motion or selfmotion is perceived. (From Dichgans and Brandt 1978.)

and Brandt 1978). Using optokinetic patterns equal


in area, we demonstrated that the peripheral retina
dominates visually induced vection, whereas central
vision (central visual field up to 30 in diameter)
dominates pattern recognition and egocentric
object-motion detection. When confiicting central
and peripheral, or foreground and background,
optokinetic stimuli are simultaneously presented,
dynamic visual spatial orientation relies mainly on
the information from the seen periphery, both the
retinal and the depth periphery. When an individual
is riding in a car at constant velo city, self-motion
sensation is produced by a relative backward motion
of the surroundings, but the driver looking in the
rearview-mirror is simultaneously able to detect and
pursue single objects moving with respect to himself
and in relation to the environment. One can demonstrate that vection is basically independent of the
direction of eye movements, if one dissociates eye
movements in circularvection by presenting a small
pattern in the centre which moves in a direction
opposite to that of a large surrounding pattern, subtending the rest of the subject's visual field (Fig.
29.3). Velo city and luminance thresholds for opto-

kinetically induced vections are elose to those for


image-motion detection, even in poor scotopic
vision (Berthoz et a1. 1975; Leibowitz et a1. 1979).
During long exposure to seen motion of the
entire visual surround, one may observe a slow
tonic decrease in perceived velo city of circularvection, and, after several minutes, periodic reversals of
the direction of perceived self-motion, concurrent
with a shift in average eye position ("Schlagfeld")
toward the slow instead of the quick phase of optokinetic nystagmus (Brandt et a1. 1974a). Periodically
reversed vertical vection has been experienced by
an Austrian cosmonaut post fiight (Mueller et a1.
1994). This indicates transient microgravity modulations of dynamic spatial orientation. Perceptual as
weIl as oculomotor aftereffects (optokinetic
afternystagmus) induced by rotatory motion of the
entire surround have shown that duration and
intensity of both phases (initial positive followed by
negative) depend on the stimulus duration (Brandt
et a1. 1974a). Whereas the positive phase increased
with increasing stimulus duration only up to 1 min,
the negative component was found to still increase
with the longest stimulus duration tested (15 min).

Visual vertigo: visual control of motion and balance

413

..

Standard 60' Is

I Fixation

Ci rcularveclion (Ve~

Optokinetic Stimulus (peripheral)


Optokinetic Stimulus (30 central)
Optoklnetic Nystagmus
Circularvection

without central field


0/
OKN

__

------

Velodty peripheral Sti mulus

Central Optokinetic Stimulation in


opposite direction

cv

__

OKN _ _

=10

10
Magnitude Estimation

15

without central fjeld


0/
OKN

Fig.29.3. Optokinetic stimulation of centre and periphery in opposite directions by means of centrally located mirror.ln this situation circularvection (CV) is determined bya peripheral stimulus, while optokinetic nystagmus (OKN, lower graph) is guided bya central
stimulus. As soon as a central stimulus is introduced, direction of OKN reverses, whereas velocity of circularvection is only slightly
decreased (see magnitude estimation in relation to unidirectional full-field stimulation at upper right). Since eye movements are in the
direction opposite to the movement of peripheral stimulus, its velocity is overestimated. (From Brandt et al. 1973.)

Visual-vestibular interaction: functional


significance of visual and vestibular cortices
Visual and vestibular information must converge on
the same locus somewhere in the brain, where they
are integrated to provide the basis for a unitary
model of dynamic and static orientation in space. It
is not yet known which visual pathways - subcortical accessory optic tract and/or cortical striate projection - convey optokinetic information to the
central vestibular system in order to make possible
the convergence that has been found. The functional
significance of the visual cortex was demonstrated in
patients with homonymous hemianopia who neither
perceived circularvection nor exhibited a postural
destabilisation when exposed to optokinetic pattern
motion (yaw or roll) restricted to the scotoma
(Straube and Brandt 1987).
The vestibular cortex (Chap. 14) is supposed to be
a major centre for the perception of self-motion secondary to labyrinthine stimulation. Accordingly,
Friberg et al. (1985) found an increase in regional
cerebral blood ftow of the superior temporal gyrus
posterior to the auditory cortex during caloric irrigation of the extern al auditory canal in humans.

Bottini et al. (1994) were the first to identify central


vestibular projections by PET. In animal experiments (monkey), an optokinetic-vestibular convergence could be demonstrated in area 2v (Bttner and
Buettner 1978; Bttner and Henn 1981) and the
parieto-insular cortex (Grsser et al. 1982; Grsser et
al. 1990a,b). It seems reasonable, therefore, to question whether the vestibular cortex is involved in the
perception of circularvection. The functional significance of the vestibular cortex as well as ipsilateral
visual-vestibular interaction was demonstrated in
patients with unilateral tumour lesions involving the
vestibular cortex areas. They either failed to perceive
circularvection or showed a significant increase of
circularvection latencies when monocular optokinetic stimulation was restricted to the visual
cortex ipsilateral to the tumour (Straube and Brandt
1987).
Visual-vestibular convergence has been demonstrated for the vestibular nuclei, the thalamus, and
the vestibular cortex, but it has not yet been established which pathways are used in which order to
mediate the perception of direction and speed of circularvection. The perception of speed is probably
insufficiently coded by the second-order vestibular

Vertigo

414

nudei neurons, because in the goldfish (Dichgans et


al 1973) and the monkey (Bttner and Renn 1981)
there is a poor correlation of frequency modulation
of these neurons with the actual speed of pattern
motion. A "speedometer function" is more adequately
provided by the vestibular cortical neurons type 1
and 2 in the area 2v, which show a good correlation
with the visual input (Bttner and Renn 1981). On
the basis of our animal experiments with frogs and
pigeons (Straube et al. 1987), as well as our data in
patients (Straube and Brandt 1987), we hypothesised
that circularvection is induced by visual motion
stimulation of the primary visual cortex, which then
activates the second-order units by descending corticopontine pathways to the vestibular nudeL They
in turn inform the vestibular cortex that self-motion
is involved. Such pathways have been demonstrated
for eye movements: the visual information first goes
forward to parietal and frontal cortex areas and
from there to the dorsolateral pontine nudei or the
superior colliculus. There are also projections from
the visual cortex to the nudeus of the optic tract,
which proceed from there and also from the superior
colliculus (via the cerebellum) to reach the vestibular nudeL One could argue that the vestibular nudei
serve as a "switch" for the perceptual interpretation
of seen motion: "on" means self-motion, "off" me ans
object-motion (Straube and Brandt 1987).
Another possible explanation is that visual
motion information is first transferred by intercortical connections from the visual cortex (area VI) to
the middle temporal and posterior parietal cortex
near the vestibular cortex areas. There, specific
neurons have been found which respond to various
parameters of pattern motion. Furthermore,
neurons in the areas MT/MST are able to differentiate between retinal shifts caused by eye movements
and those caused by pattern motion. Descending
pathways from the vestibular cortex to the vestibular
nudei have been indirectly demonstrated. In addition,Arslan and Molinari (1965) and Gildenberg and
Rassler (1971) found that stimulation of the temporallobe produced a short latency modification of
the activity of the vestibular nudeL
The neurophysiological correlate of the different
characteristics of the visual stimulus (such as area of
the visual field), which determines whether pattern
motion is perceived as object-motion or as selfmotion, is still not known. It is undear if the decision whether circularvection is perceived or not
depends on the activation of the vestibular nuclei,
which send information about onset as well as
direction of self-motion to the cortex via the
vestibulothalamocortical projections. The finding,
however, that fixation suppression of optokinetic
nystagmus also suppresses normal modulation of

vestibular nudei neurons in the alert monkey


(Bttner and Buettner 1978; Bttner and Renn 1981)
does not support this interpretation. It is weH known
from psychophysics in humans that stationary fixation does not affect circularvection. Thus, visualvestibular convergence in the vestibular nuclei may
possibly be independent of the perception of selfmotion and may reflect only eye position, namely
deviation of the "Schlagfeld". At least the velo city of
self-motion is mediated by direct interaction
between the visual and the vestibular cortices. The
fact that circularvection latencies rather than perceived velo city were affected in patients with tumour
lesions of the vestibular cortex lends support to the
argument that ipsilateral visual-vestibular cortex
interaction is also involved in the decision to elicit
circularvection.
To determine the unknown visual-vestibular
interaction during circularvection, we conducted a
PET activation study on circularvection in human
volunteers (Brandt et al. 1998). The PET images of
activated cortical areas during random dot motion
stimulation without circularvection were compared
to those with coherent roll motion stimulation that
induced vection. Activation of the medial parietooccipital cortex, separate from the well-known
motion-sensitive areas MT/MST, is a positive (activated) correlate ofcircularvection (Fig. 29.4). A significant positive correlation was shown between
perceived intensity of circularvection and the relative rCBF increases in this area. This area corresponds best to a visual area in the parieto-occipital
fissure, which was activated in a PET study that compared observation of a moving random dot with that
of a stationary dot pattern (Dupont et al. 1994).
Contrary to our expectation, it was also shown that
visual motion stimulation with vection not only activates this medial parieto-occipital visual area, but
simultaneously deactivates the parieto-insular
vestibular cortex (Fig. 13.3; Brandt et al. 1998). This
supports the hypothesis of a reciprocal inhibitory
visual-vestibular interaction as a basic sensorimotor
mechanism for adequate self-motion perception (see
Chap. 13). In an earlier PET study on the effects of
caloric vestibular stimulation, we showed activation
of the vestibular cortex with concurrent deactivation
ofthe visual cortex (Wenzel et al. 1996).

Rollvection-tilt: optokinetic graviceptive


mismatch
Gravitational orientation depends on two major
sources of experience: the exteroceptive visual and
the proprioceptive postural. A stationary observer,
while viewing a large visual scene rotating around

Visual vertigo: visual control of motion and balance

415

?{p

'N

40 / s
lateral postural sway

CON

ON

20 Nm
ccw

Fig.29.4. PET activation study on rollvection. Comparison of


the relative rCBF increase under visual motion stimulation, wh ich
induced rollvection (counterclockwise); control condition did not
induce rollvection (random dot movement). All significant voxels
above the statistical threshold (p > 0.001, corrected for multiple
comparisons) are shown. The medial views and the transversal
images illustrate the selective bilateral activation of areas at the
border between occipital and parietal cortex area (top).
Correlation analysis illustrating the voxels with significant, positive correlation to perceived intensity of rollvection (p < 0.05
after Fisher's transformation from r to z). All significant voxels
above the statistical threshold are displayed. There is a positive
correlation with the perceived intensity of rollvection in an area
at the border between occipital and parietal cortex area and of
rostral anterior cingulate (bottom). (From Brandt et al. 1998.)

his line of sight, experiences a continuous sensation


of self-motion opposite in direction to that of the
pattern motion (rollvection; Fig. 29.5) but only a
limited body tilt and dis placement of visual vertical
(Dichgans et al. 1972). Induced displacement can be
measured in terms of tilt angles of the apparent
visual and postural vertical from gravitational vertical and may be conceptualised as the result of a compromise that weighs the different and, in part,
contradictory sensory inputs for gravitational orientation. Tilt angles for pitch are smaller than those for
roll, and in addition exhibit an asymmetry between
the smaller pitch-up and larger pitch-down angles
(Young et al. 1975). They appear to be fixed with
respect to the body (egocentric localisation), rather

J--w.-1

Fig.29.5. Clockwise (CW) rotation of the visual scene causes a


counterclockwise (CCW) dis placement of perceived visual and
body vertical (2). Visuospinal postural compensation for this
apparent displacement causes a measurable body tilt in the
direction of pattern motion (3). Original recording of lateral body
sway (bottom) in humans with a tonic body shift after some seconds of latency and an aftereffect following the stimulation
period.

than with respect to the direction of down (absolute


localisation). A similar asymmetry is observed for
postural reactions to a linearvection stimulus, with
the tendency to fall backward considerably stronger
than the tendency to fall forward (Lestienne et al.
1977).
The apparent displacement of both visual and
postural vertical has been explained as the result of a
central recomputation of the orientation of the gravity vector based on vision. This is supported by the
finding that the visually induced tilt markedly
increases when the otoliths are placed in a less
favourable position by lateral head tilt (Dichgans et
al. 1974; Young et al. 1975). The potentiation of the
net visual effect on the apparent vertical in an
inclined head position is maximal if the visual stimulus is moving opposite to the head tilt. This asymmetry can be interpreted as the functional
hypertrophy of a biologically adaptive mechanism
that assists postural righting when the visually
induced tilt perception adds to the otoliths' information. Under natural conditions body tilt causes relative motion of the visual scene in the direction
opposite to that of body displacement. The induced
relative motion of the visual environment will, in

Vertigo

416

this case, correlate with the actual information from


graviceptors for postural adjustment. Psychophysical
adjustments of perceived "horizontal" (eye level)
during combined changes of gravitoinertial force
level and visual field pitch were consistent with a
model in which visually perceived eye level is treated
as a consequence of an algebraically weighted average or a vector sum of visual and graviceptive inftuences (DiZio et al. 1997).
Patients with profound bilateral vestibular loss
experienced complete unambiguous self-rotation
with asensation of rotating fully through an upsidedown orientation when exposed to vection about the
pitch or roll axes (Cheung et al. 1989). This suggests
that in the absence of otolithic cues, static
somatosensory input alone is insufficient to counterbalance the effects of visual pattern stimulation.

Visual pseudo-Coriolis effect and


pseudo-Purkinje effect
Cross-coupled accelerations (Groen 1961) occur
when the head, while being rotated about one axis, is

tilted about a second axis. The resulting extreme


spatial disorientation is called the Coriolis effect
(Schubert 1931) and is of particular concern to airplane pilots.
Pseudo-Coriolis effects and optokinetic motion
sickness are elicited by bending the head out of the
axis of rotation of a circular visual surround, when
that moving surround induces the illusion of selfrotation (Fig. 29.6). With respect to tilt sensation and
vegetative symptoms, optokinetic pseudo-Coriolis
effects correspond to vestibular Coriolis effects
(p. 488) that arise from similar head movements
when the body is actually rotating (Brandt et al.
1971). Quantitatively, pseudo-Coriolis effects are
smaller in magnitude than Coriolis effects, and saturate at lower velocities of stimulation (90-100 0 /s).
Pseudo-Coriolis effects depend on the illusory sensation of self-rotation (circularvection) and share its
prolonged time course after stimulus onset and termination, its relation to stimulus velocities, and its
dependency on the stimulus area in which the retinal periphery predominates. Optokinetic inftuence
on vestibular Coriolis effects depends on the direction and speed of the moving visual stimulus and
results in either inhibition or facilitation of apparent
tilt and nausea (Brandt et al. 1971).

Vestibular Co riolis effect

Visual pseudo-Corio lis effect

Fig.29.6. Coriolis effects arise from cross-coupled accelerations when the head, while undergoing a rotation of constant angular
velocity, is tilted about an axis perpendicular to the axis of rotation. Optokinetic pseudo-Coriolis effects are elicited in objectively
stationary subjects by bending the head out of the axis of rotation of circular visual surroundings; the moving surroundings induce
the illusion of self-rotation, circularvection. (From Brandt 1984.)

Visual vertigo: visual control of motion and balance

An active head tilt during a post-rotational semicircular canal response causes a tumbling sensation
of turning about a tilted axis, with corresponding
postural instability and slight nausea (Purkinje
1820). The angular velocity vector located relative to
the skull moves with the head, and the sensation of
an off-vertical rotation confticts with graviceptive
information deriving from the otoliths. A comparable but visually induced phenomenon can be
observed if a subject is simultaneously exposed to a
rotation of the visual surround about the line of
sight (horizontal axis) during angular acceleration
about an earth vertical axis (head aligned to the
z-axis) . Optokinetic stimulation then causes an
apparent head and body tilt opposite to the pattern
motion (rollvection), which leads to an unpleasant
perception of rotation around an off-vertical axis
(Fig. 29.7; Brandt 1984).
Both the vestibular and the optokinetically simulated "Purkinje effects" are related to Coriolis effects.
The semicircular canals in visual pseudo-Coriolis
effects, and the otoliths in visual pseudo-Purkinje
effects are simulated optokinetically as a result of a
visual-vestibular convergence.

417

Optokinetic motion sickness


It is an old observation that "vestibular" motion

sickness does not occur in individuals who lack a


functional vestibular apparatus (James 1882). Since
blind persons are susceptible to acute motion sickness (Desnoes 1926) and exhibit no significant difference in the incidence rate of motion sickness
from that of a control group of normal subjects
(GraybieI1970), it was concluded that vision is not
an essential but rather a secondary aetiological factor in the genesis of motion sickness during passive
transportation. On the other hand, when bilaterally
labyrinthine defective patients were exposed to visual
roll or pitch stimulus patterns, no motion sickness
was observed (Cheung et al. 1991).
In the absence of actual body movements, however,
pure "optokinetic" motion sickness has been reported
in healthy subjects while viewing a tilted swinging
room (Wood 1895; Witkin 1949; Lishman and Lee
1973), a rotating cylinder from inside (Crampton
and Young 1953), wide-screen movies (Money 1970;
Parker 1971), a moving scene from the screen of a

Vestibular Purkinje effect

Visual pseudo-Purkinje effect

ZAXiS~
I

x AXiSt J

Fig.29.7. Vestibular Purkinje effects are induced by an active head tilt during a post-rotational semicircular canal response; they
cause a tumbling sensation of turning about a tilted axis with corresponding postural instability and slight nausea . A comparable
"visual pseudo-Purkinje effect" can be observed if a subject is simultaneously exposed to rotation of the visual surround about the line
of sight (rollvection) during angular acceleration about an earth vertical axis. (From Brandt 1984.)

Vertigo

418

Hight simulator (Miller and Goodson 1960; Barret


and Thornton 1968; Reason and Diaz 1971), or when
exposed to visual virtual environments (Kennedy
and Stanney 1996). Optokinetic motion sickness
induced by oscillation of the surroundings has been
compared quantitatively with vestibular motion
sickness induced by oscillation of the subject in the
dark; in both cases the oscillation was about the vertical (z) axis with a frequency of 0.02 Hz and a peak
velo city of 100 0 /s. The comparison showed that the
optokinetic stimulation tended to be more effective
in eliciting motion sickness (Dichgans and Brandt
1973). This observation may be explained by the
visual-vestibular mismatch in the case of motion of
the surrounding alone. Reports on optokinetic simulator sickness (Miller and Goodson 1960) have
shown that pilots with real Hight experience (and
therefore a stored expectation for sensory input patterns due to certain manoeuvres) exhibit a high er
susceptibility than non-pilots. Motion sickness susceptibility to optokinetic stimulation correlates with
a past history of motion sickness (Hu et al. 1996). A
spatial frequency effect has been described by
Diener et al. (1976) to subserve motion constancy
both in object- and self-motion perception. A pattern of constant physical size, if placed at a further
distance, will be imaged with a higher spatial frequency on the retina than a doser pattern. The
severity of vection-induced motion sickness is also
affected by different spatial frequencies (Hu et al.
1997).
Visually induced paroxysmal nausea and vomiting have been described as a presenting manifestation of multiple sderosis (Khan et al. 1995; Bronstein
1996) and resemble central vestibular positioning
vomiting (p. 293). According to the mismatch concept (see Motion sickness, p. 487), vision has been
reported to promote motion sickness during locomotion or passive transportation, if it is at variance
with actual body accelerations, whereas it is suppressed if vision is in physiological agreement with
the simultaneous vestibular input (Brandt et al. 1971;
Dichgans and Brandt 1973).
Anti-motion sickness drugs, such as the belladonna
alkaloid scopolamine (0.6 mg) or the antihistamine
dimenhydrinate (100 mg) effectively prevent both
vestibular and optokinetic motion sickness (Brandt
et al. 1974b). However, they simultaneously cause
remarkable central sedative side effects and decrease
performance, as documented by psychological
efficiency tests.

Physiological height vertigo and


postural balance
Height vertigo, a visually induced syndrome commonly experienced on top of high structures, is
manifested by a subjective instability of posture and
locomotion, coupled with a fear of falling and vegetative symptoms, which spontaneously remit after
the situation that induced the stimulus is terminated.
Habituation to height vertigo may occur through
repeated exposure, as is observed in steeplejacks,
roof-workers, and tight-rope artists, who achieve a
remarkable degree of postural balance with seeming
insensitivity to height. The textbooks generally
attribute height vertigo to psychopathological
processes such as neurotic acrophobia (see p. 460),
and psychoanalytical attempts have been made to
explain its pathogenesis (Baumeyer 1953; Rennert
1990). Surprisingly enough, it is difficult to find any
controlled experimental study in the literature on
this distressing phenomenon, which is not only
"known" to animals (see Visual diff behaviour,
p. 422) but is part of the personal experience of nearly
every human being. Erasmus Darwin's approach to
the problem in 1794, however, still appeals because
its concept of multisensory interaction as weIl as the
relation between postural imbalance and vertigo is
still topical:
"Anyone, who stands alone on the top of a high
tower, if he has not been accustomed to balance himself by objects placed at such distances and with
such indinations, begins to stagger, and endeavours
to recover hirns elf by his muscular feelings. During
this time the apparent motion of objects at a distance below hirn is very great and the impressions of
this apparent motion continue a litde time after he
has experienced them; and he is persuaded to incline
the contrary way to counteract their effects; and
either immediately falls, or applying his hands to the
building, uses his muscular feeling to preserve his
perpendicular attitude, contrary to the erroneous
persuasions of eyes!'
There is a simple geometrical explanation for
physiological height vertigo being a "distance vertigo": postural balance is destabilised when the distance between the observer's eye and the nearest
visible stationary contrast becomes critically large
(Brandt et al. 1980; Bles et al. 1980). Thus, a visually
induced postural imbalance causes a "physiological
vertigo syndrome." This syndrome can dearly be
contaminated with additional cognitive and psychological factors that are mainly responsible for height
anxiety and the vegetative symptoms.

Visual vertigo: visual control of motion and balance

To maintain postural stability in the upright position, afferent signals must be generated as an input
for motor compensation of natural fore-aft and lateral body sway; head sway is horizontal tilt rather
than roll tilt over the first several centimetres (parallel
shift) because of mechanical parallel motion of the
pelvis. Optimal balance requires continuous central
evaluation of the re afferent sensory consequences of
self-generated body movements. Vision plays a
major role in postural stabilisation (see p. 423). A lateral head or body sway, for example, causes either a
shift of the visual surround on the retina when the
eyes are stationary in the head or an eye movement
of the same angle if the object is fixated, involving
afferent and efferent motion perception, respectively.
Simple geometrical analysis indicates that body
sway must increase with increasing distance
between the eyes and the nearest stationary contrasts in order to be detected visually (Fig. 29.8). The
greater the distance y is to the object, the smaller is
the angular displacement a on the retina. Since the
natural lateral head sway is about 2 cm in amplitude,
it is important to determine the specific eye-object
distance at which this sway amplitude could not be
visually detected. Since an angular displacement of
the visual scene of 20 min of arc is necessary for
detection by the paracentral and peripheral parts of
the retina (Aubert 1886; Leibowitz 1955), anormal
lateral head sway of 2 cm would be the subthreshold
at a distance of ab out 3 m. This leads to a perceptual
conftict, since the vestibular and somatosensory
receptors sense a body shift that the visual system
cannot detect. There are two possible explanations
for the consequent increase in sway amplitude. The
simplest explanation is that the normally primary
visual input suppresses the sensory conftict, and the
increased sway amplitude reftects the inability of the
visual system to detect and correct small sway movements: the system oscillates between points where
visual stimulation is above threshold. Alternatively,
the posture control system may try to resolve the
conftict by actively generating larger sway movements with the intention of producing suprathreshold visual stimuli. For a simple single loop control,
one could expect a relationship between distance
and sway amplitude (tan a = x/y), with the gain
dependent on the retinal movement detection
threshold. However, with the lower limbs dose
together in free stance, a person falls over at a head
sway exceeding 10 cm. Thus, at an eye-object distance of 15-20 m, a "maximal body sway" would produce retinal angular deviations (a) less than 2 min
of arc according to the trigonometrical model. Since
we are dealing with a multiloop control of postural
balance, it can be assumed that with increasing sway
amplitudes, the particular sensory weight of the

419

somatosensory and vestibular afferences becomes


greater, increasing their contribution to postural
stabilisation.
The trigonometrical model developed for afferent
motion perception during lateral horizontal body
sway must also be valid for efferent motion perception, since the thresholds of movement detection are
about the same with eye pursuit and widely
"stabilised retinal images". It must also hold for foreaft sway in which the retinal images vary in size with
the sway (Fig. 29.9). Involvement of eye torsion - in
which the retinal shift is independent of the distance
between eye and object - can have consequences for
posture only in the case of great lateral (rotational)
sway amplitudes.
It seems reasonable to condude that the displacement angle on the retina, dependent as it is upon
distance, cannot be a determinant of the net compensatory sway. Higher-order depth constancy
mechanisms, similar to those for size and motion
constancy, must be involved in compensating for the
retinal deficiency. This is evident in normal structured surroundings when the body sways in front of
several stationary objects at different distances. Here
monocular and binocular depth cues, as well as
motion parallax, may be used to stabilise posture.
The gain of the motor reaction to a visually sensed
sway requires a precise multisensory calibration
based on all mechanisms subserving self-motion
perception. A retinal shift of about 20 min of arc
seems to be the minimal requirement for the visual
afferent contribution.
To validate this hypothesis, aseries of psychophysical and posturograpic experiments were
performed in the laboratory and under natural conditions on high buildings (Brandt et al. 1980; Bles et
al 1980). The results were consistent with the
trigonometrical model:

The occurrence of physiological height vertigo


is clearly related to body position. It is strongest
during erect stance, when maintaining balance
is most difficult, and minimal when the subject
is lying down (Fig. 29.10);
height vertigo can be induced either by downward or upward gaze, because it is the distance
rather than the visual direction that is critical;
height vertigo appears at a distance of about
3 m, increases with increasing altitude but saturates at less than 20 m; thus, vertigo is already
maximal at a net altitude of about 20 m;
lateral and fore-aft body sway (posturographic
experiments) as well as head sway increase nonlinearly with increasing eye-object distance;
nearby stationary contrasts in the periphery of
the visual field alleviate both subjective vertigo
and measurable postural imbalance.

420

Vertigo

W
7

200

150

J!
=ti

..

.
~

100

GI

....
\

\
\

.f

"I:

CL

"'

Oll

CL

50

-!

tt

GI
U

\
\

CL

&;

..$ '"

l!
.5

'"

&:

GI

\
\

........-'\

ClI!JEC'T

GI

&:

\
\

\
\

0
lana

10

Lateral hlld dlspllcement 10m)

..L
y

I
I
I
I
I
I
I
I

200 m

I
I

I
I
I
I
I

Eyes closed

1
1
1

fore - aft sway

10Nm ]~~v
10Nm ]

laie ral s w a y :

J"\..

\J~
I

h.~ _~

1
1
1

I
I
I

1-

......,..

\.J"Vv1~

Fig.29.8. Mechanism of physiological height vertigo. Geometrical analysis indicates that, in order to be visually detected, body sway
must increase with increasing distance between the eyes and the nearest stationary contrasts.Angular displacements (X,on the retina,
caused by a lateral head displacement, are smaller, the greater the distance y is to the object. Diagram shows relationship between
head-object distance, y, and lateral head displacement, x, for given retinal dis placement threshold of either 2 or 20 minutes of are.
However, because postural regulation involves the multiloop control, additional proprioceptive cu es may alter sway amplitude as weil.
Fore-aft and lateral body sway (original traces) with eyes closed, eyes open in front of a wall, and eyes open on a high building with
and without additional stationary contours in the peripheral visual field. Sway amplitudes increase under height vertigo conditions,
especially in the low frequency range. Simultaneously, nearby stationary contrasts in the seen periphery stabilise "height vertigo
sway." (From Brandt et al. 1980.)

Visual vertigo: visual control of motion and balance

421
EYES
CLOSED

n" 6
60
ID

50

OBJECT

1
f

FORE-AFT

SWAY

60

10

15

20

25

75

(m)

EYES
CLOSED

n" 6

50
40

LATERAL SWAY

''' I

30

" ,

t
o

10

15

20

25

15

m)

Fig.29.9. Differential effects of lateral and fore-aft head sway on retinal image shift of a viewed stationary object (feft). Body sway
during upright stance as a function of the distance between the head and the seen environment (fight). Fore-aft (top) and lateral (bottam) body sway (mean root mean square va lues, RMS) at different eye-object distances (0.5,2.5,5,10,15,20,25, and 75 m) both without (*l. with single (e), and with double (. .. ) slice of foam rubber on top of the posturography platform. Body sway increases with
increasing eye-object distances. Visual stabilisation of posture is reduced under stimulus conditions that elicit height vertigo.

The significant increase in body sway amplitudes


introduces areal danger of falling from a high place.
Thus, physiological height vertigo is a meaningful
warning signal to the body to withdraw from stimulus situations that cannot be adequately perceived in
terms of the space constancy necessary for postural
control. The "static" stimulus condition of height or
distance will not, per se, destabilise posture and
result in an irresistible fall because of the redundancy
in the controlloops. The lack of an appropriate input
is broadly compensated for by the somatoreceptors
and the labyrinths. However, additional disturbances
such as wind or an unstable foot support may result
in serious problems in maintaining an upright position, since subjects are apt to use the "false"
visual information in problematic circumstances.
Therefore, patients with vestibular or somatosensory
dysfunctions (e.g. polyneuropathy) are subject to a
particulady great risk when exposed to height vertigo

situations. Distance vertigo does not occur on the


ground, despite the fact that visual targets may be at
great distance, because cues from nearby stationary
contrasts (provided by peripheral vision) prevent
the causative sensory mismatch. It might, however,
occur if the ground is not aligned to the earth horizontal.
For a given altitude, subjective height vertigo can
be induced at an angle of decline of 40-50 if there is
a progressive increase with increasing slope. A maximal vertigo is reached at about 70-80.
With increasing slope, peripheral vision provides
less information about nearby stationary contours.
According to posturographic measurements, the
development of subjective height vertigo also
requires an exposure time of 5-30 s for saturation
conditions to be reached.
Thus, there is a "visuospinal" mechanism that
explains physiological postural imbalance and vertigo

Vertigo

422

Staring at moving objects such as clouds


increases the danger of falling, because visually
induced vection causes additional postural
destabilisation.
One should avoid long exposure times, because
height vertigo usually takes several seconds to
develop.
Looking through binoculars is very dangerous,
because it restricts the visual field and introduces an unusual, and therefore unadapted,
magnification factor.
Body and head position should be adjusted to
the gravitational vector because vision will
receive a relatively greater sensorial weight
(which is undesirable) if the otoliths are displaced beyond their optimal working range by
extreme head tilt (see Physiological head extension vertigo, p. 297). Other observations suggest
that the feet should be firmly plan ted on an
"earth horizontal" surface.

Licence for workers at heights?

Fig. 29.10. The magnitude of subjective height vertigo is


related to body position. It is greatest during upright stance,
when maintaining postural balance is comparatively most
difficult.

at heights, which is separate and distinct from the


purely cognitive or psychological factors stressed by
earlier researchers such as Purkinje (Purkinje 1820;
Kobrak 1924).

Physical prevention of physiological height


vertigo
Knowledge of the stimulus characteristics required
for the optimal visual contribution to stabilise postural balance also dictates practical advice to reduce
height vertigo in susceptible subjects (Brandt et al.
1980):

One should avoid the free upright stance in critical situations at high altitudes. This is done
intuitively when grasping for a stationary
framework or leaning against a wall for support.
When looking down, one should obtain stationary cues from nearby contrasts in the peripheral
visual field. Visual stabilisation of posture is
then maintained by the retinal periphery, while
the central retina mainly serves egocentric
recognition and pursuit of objects.

The professional qualification of men working above


ground level (roof-workers, steeplejacks, etc.)
depends on optimal sensorimotor control because of
the continuous risk of a catastrophic fall. As in the
case of driving licences, generallimitations are given
by diseases that possibly affect attention or consciousness, such as epilepsy, narcolepsia, AdamsStokes or drop attacks (p. 389), malign arterial
hypertension, addiction to drugs or alcohol, etc.
Specific limitations are provided by either neurotic
acrophobia (to be evaluated by psychiatrists) or any
deficiency in the multiloop control of postural balance (to be evaluated by otoneurologists and/or
ophthalmologists). Although
the
potential
decrement in balance is of minor relevance at
ground level (due to e.g. external eye muscle paresis,
polyneuropathy, benign paroxysmal positioning
vertigo, cerebellar ataxia, parkinsonism), it may be
responsible for a headlong fall due to the reduced
and sometimes misleading sensory input when
working at heights. In case of doubt, a specialist with
clinical experience in posturography should be consulted, since standardised tests or normal values for
licensed work at heights are not yet available. The
simple testing of vestibular function, however, is
one-sided and insufficient; it does not correlate with
sensitivity to vertigo (Gramowski 1962).

The "visual cliff" phenomenon


The visual mechanism underlying physiological postural imbalance at high altitudes is possibly related to

Visual vertigo: visual control of motion and balance

a broadly gene-linked depth avoidance, known as


visual cliffbehaviour (Fig. 29.11). This phenomenon
has been comprehensively studied by Walk and
Gibson (Walk et al. 1957; Walk and Gibson 1961).
The innate ability of the human infant to visually
perceive and avoid a brink is supported by animal
experiments with a number of species including
chicks, rats, kittens, and goats. Visual depth avoidan ce does not require prior experience of faHing off
places; it can even be observed when locomotion
begins, at birth. In an ambivalent stimulus situation
in which a heavy glass covered the deep side of a
cliff, crawling children still refused to cross the glasscovered brink despite the somatosensory tactual
assurance of asolid surface (Walk and Gibson 1961).
There are two visual cues for perceiving height as a
special case of distance: (1) the drop-off increases in
texture density beyond the edge and (2) absolute
and relative motion paraHax cues differ due to active
locomotion and head movements. Both texturel
density preferences (De Hardt 1969; Davidson and
Whitson 1973) and motion paraHax cu es are utilised
and interact. Although Doris De Hardt (1969) was
able to demonstrate in rats that pattern size may
override confticting information from relative
retinal motion, Walk and Walters (1974) still believe
that motion paraHax seems to be a major determinant of depth perception, as long as adequate text ure
is provided from aH portions of the visual scene.

Vision and posture


Vision plays an important but not essential role
within the redundant multisensory control of postural
balance in a gravitational field. Physiologically it
attenuates self-generated body sway by 50% (Travis
1945; Edwards 1946). Clinically its effect on postural
stability accounts for the significance of the
Romberg test (Romberg 1846) for patients suffering
from either proprioceptive sensory dysfunction or
cerebellar ataxia. Under ordinary conditions it is
possible to maintain upright posture with the eyes
closed. However, the visual contribution to postural
regulation becomes dominant in patients with
defects of the vestibular or somatosensory system, as
weH as in normal subjects when performing more
demanding balancing tasks, such as those involved
in sports or riding a bicycle.
Optimal balance requires the continuous monitoring of sensory afferents generated by body sway. The
decisive visual signal that starts active correction of
body tilt may depend on either the relative shift on
the retina of the image of stationary visual sur-

423

Fig. 29.11. "Visual eliff" phenomenon. Animal experiments


provide support for the human infant's innate ability to visually
perceive and avoid a brink.

roundings or "efferent motion perception" when fixating stationary targets during head motion. The
weight of visual input to the multisensory control
process increases in the elderly (see Chap. 27), a
finding that is supported by the increase of the
Romberg quotient (body sway with eyes closed/eyes
open) with increasing age (Straube et al. 1989).
While the amplitude, duration, and direction of
body acceleration clearly define a vestibular stimulus, the variability of visual motion stimuli gives rise
to greater complexity. The visual perception of relative motion due to head sway is determined by the
three-dimensional structure of the environment as
weH as the viewing conditions, for example, the level
of illumination or degree of accommodation.
Additional factors such as visual acuity, the location
and size of the stimulus within the visual field, and
changing eye-target distances (and, therefore, different retinal velocities) can all be involved in varying
combinations.
The following argument demonstrates how visual
performance and differential characteristics of the
visual stimuli may significantly affect optimal postural balance. In agreement with Droulez et al.
(1985) and Nashner (1985), we believe that visual
input is not only used for continuous evaluation of
head sway (feedback sensorimotor loop) but also for
discontinuous adjustment or as a trigger for programmed postural strategies, i.e. patterns of muscle
activation. The synergies of muscle activation, programmed by prior experience of active stance and
locomotion, may be guided by sensory input from
the joints, the muscles, the labyrinths, or the eyes,
individually or in combination. It is unlikely that
there is a single dominant relationship between visual

424

stimuli and postural responses, which arises from a


specific visuospinal controlloop. On the contrary,
the evidence suggests that the posture control system
is able to utilise the accuracy and low threshold for
movement detection of the visual system in a way
appropriate to the circumstances by recalibrating or
reorganising the control configuration. The use of
visual feedback to further decrease body oscillations
(Gantchev et al. 1985) supports this assumption.

Moving visual scenes


Under natural conditions (e.g. looking at moving
clouds) and in the laboratory, vision may have a
strong destabilising effect on posture, particularly
when visually perceived motion does not adequately
correspond to the actual body shift sensed by the
vestibular and somatosensory systems. Observations
of optokinetically induced body sway were reported
in the older literature (Fischer and Kornmller
1930). This phenomenon has been extensively investigated using circular motion of artificial surroundings about the vertical z-axis (Kapteyn and Bles
1977), roll motion about the line of sight (Dichgans
et al. 1976), a sinusoidally tilting room (Bles et al.
1977), and linear motion of the seen environment
(Berthoz et al. 1975; Lee and Aronson 1974). In these
instances there was an optokinetically induced illusory perception of self-motion that was opposite in
direction to pattern motion (see Dichgans and
Brandt 1978 for a review of this subject). The postural
"reflexes" compensate for this subjective body shift,
resulting in a measurable body tilt in the direction of
pattern motion. These optokinetically induced postural reactions require stimulation of the entire visual fie1d, or substantial portions of the periphery,
which corresponds to the stimulus characteristics
for visually induced vections (p. 412). Postural control in patients with vestibular disorders was particularly affected by optic flow stimuli (Redfern and
Furman 1994).

Visual acuity
Normal or only slightly reduced visual acuity is necessary for tasks such as reading or pattern recognition involving perception of high spatial
frequencies. Postural control is less sensitive: a logarithmic decrease in visual acuity pro duces a linear
increase in postural instability as measured by"root
mean square" (RMS) sway amplitude. This effect is a
result of the impaired perception of the motion of
the head relative to the surroundings (Paulus et al.
1984; Brandt et al. 1985).

Vertigo
Visual acuity can be systematically diminished in
controlled steps by semitransparent plastic foils,
normally used to penalise the good eye in children
with squint amblyopia. The foils, attached close to
both eyes, produce a whole fie1d reduction in visual
acuity of 0.6, 0.3, 0.1, 0.03, and 0.01. Visual acuity of 1
is defined by the detectability of a gap of 1 minute of
arc in a Landolt C ring which, at a distance of 6 m,
corresponds to a gap of 1.75 mm. With an acuity of
0.1, this gap can be detected at 60 cm and with an
acuity of 0.01, at 6 cm.
The increase in postural sway is significant for all
acuity steps tested (Fig. 29.12), but quantitatively
moderate with slight reduction for visual acuities of
0.6 and 0.3. For subjects standing on one layer of
foam rubber (to disproportionately enhance the sensorial weighting of vision), the visual contribution to
postural control ceases when the "Ganzfe1d" reduction in visual acuity is lower than 0.01 for lateral and
0.03 for fore-aft body sway, because sway under
these conditions is the same as that with eyes closed
(Paulus et al. 1984). Visual blurring by foils or glasses
impairs not only postural balance but also objectand self-motion perception (Straube et al. 1990). Use
of vision for spatial orientation during walking is
not as severely affected as the postural data may suggest. Subjects feel unstable during locomotion if
visual acuity is diminished by 99%, but they are usuaUy still able to completely avoid major obstacles in
the environment.
Patients with decreased visual acuity frequently
complain about postural instability, particularly in
the early stages. Incremental changes of visual acuity
occur in patients with dioptric abnormalities who
alternately wear or do not wear glasses, and a significant instability can be demonstrated in myopic
patients with glasses off (Fig. 29.13). Although these
persons are subjectively unaware of induced postural
instability, the normal ratio between body sway with
the eyes open compared with the eyes closed
(Romberg quotient, RMS amplitudes) shrinks significantly. Accommodation, with its corresponding
alteration in visual acuity, may playa significant role
in fine-tuning postural balance. Restriction of the
accommodation range with blurred vision in the
grasping space occurs in normal subjects older than
40 years of age, owing to presbyopia. Patients with
aphakia following cataract surgery without refractive correction in the acute stage were demonstrated
to have no measurable visual stabilisation of postural
sway when their results were compared with results
with the eyes closed (Brandt et al. 1982). This is consistent with the above results in normal subjects,
since defocus in aphakic patients causes a visual
acuity of less than 0.01. FinaUy, as already noted by
Adler in 1941, "individuals who wear strong lenses

425

Visual vertigo: visual control of motion and balance

RMS (%)

500
>-

400

I'G

fore - aft

VI

...

300

200

C;

lateral

:::I
VI

c..

100

-~.~_

.. --'

eyes closed

1.0

0.6

0.3

0.1

0.03

0.01

visual acuity
Fig.29.12. Fore-aft and lateral body sway activity as a function of visual acuity (1.0-0.01) as a percentage ofthe standard (100% raot
mean square, RMS values) at a visual acuity of 1.0. The subjects stood freely on one layer of foam rubber to enhance the sensory
weight of vision disproportionately,according to the multisensory mismatch concept. A logarithmically decreasing visual acuity causes a linearly increasing postural instability; visual stabilisation ceases at 0.01, when body sway is not significantly different from the
eyes-closed condition. (From Paulus et al. 1984.)

for the first time, particularly after cataract operations, also suffer from the strong prismatic effect
produced from looking through the peripheral parts
of the lens" and, according to our posturographic
measurements, exhibit an increased body sway
(destabilisation) similar to those with bifocal or trifocal glasses. Contact lenses provide much-improved
perception of space constancy during locomotion
and postural regulation. Practice and training of
patients with acute abnormalities of refraction may
promote sensory rearrangement so as to diminish
their spatial disorientation and postural destabilisation (Brandt et al. 1986).

Near vision and eye-object distance


Visual stabilisation in the grasping space (10-100 cm)
involves eye-object distance as a critical physical
stimulus characteristic. For geometrie reasons the
relative retinal shift of a viewed scene is greater, the
nearer the objects are to the eyes. The question is
whether or not the gradual approach of a subject to a
stationary surround within the grasping space is
reflected in a gradual improvement of postural balance. Bles et al. (1980), who were interested in gross
differences between near and far vision (Fig. 29.9),
demonstrated such a relationship earlier. Under con-

ditions of near vision, convergence is involved in


addition to accommodation, and may playa significant role in fine-tuning postural balance. Both produce double vision of visual targets before or beyond
the plane of fixation, with a corresponding alteration
of visual acuity.
The gradual approach of a subject to a flat stationary background produces a gradual improvement of
postural balance. This relationship between decreasing sway activity and decreasing eye-object distance
within the grasping space corresponds to the net
retinal displacement of the viewed stationary scene
and is best reflected by a linear characteristic
(Paulus et al. 1984).
The dependence of postural sway activity on critical visual stimulus characteristics must be considered in the design of clinical tests for postural
balance. The so-called Romberg quotient (postural
sway activity with the eyes open versus eyes closed),
for example, requires exact definitions of visual
acuity and eye-object distance.
Simple geometrie analysis indicates that body
sway must increase with increasing distance
between the subject's eyes and the nearest stationary
contrasts for visual detection to occur. This relationship mayaiso explain the mechanism of physiological height vertigo, which is a distance vertigo
produced by visual destabilisation of postural

Vertigo

426
... - RM S 1/a ' B3

RMS lal . 71
eyes open ..
glcsses
'3.5 dpl.

eyes open
w ilhou l
g lo sses

.... RMS fl 8 172

RM S laI

' IBI

lOS

Fig. 29.13. Original recordings of fore-aft and lateral body


sway in a myopic subject (27-year-old male) with either glasses
on (top) or off (middle) compared with eyes closed (bottam). The
incremental change in visual acuity due to glasses off impairs
postural balance (when standing on a slice of foam rubber) and is
sufficiently explained by the concurrent decrement of motion
perception.

balance when the distance between the subject's eyes


and the nearest visible stationary contrasts becomes
critically large (see Height vertigo, p. 418; Bles et aL
1980; Brandt et al. 1980).

Visual control of fore-aft versus lateral body


sway
Although both fore-aft and lateral body sway
increase with reduced visual function, amplitudes of
fore-aft sway are higher than those of lateral sway.
An explanation for this must be sought (1) in the differing anatomical configurations in the two planes
and the position of the feet, (2) in the differing eyetarget geometry, which affects motion detection, and
(3) in the differing modes of motion perception,
afferent and efferent. During fixation of a stationary
target, lateral head displacement is monitored mainly
by using information derived from retinal slip or
pursuit eye movements.
Fore-aft sway with a central target, however, can
be detected by two mechanisms: change in disparity
and change in target size (Regan and Beverly 1979).
Change in disparity is a purely binocular mechanism, with the greatest sensitivity at shorter distances.
It involves efferent motion perception for vergence
movements. Changing size of a visual target by a

change in eye-object distance is a purely sensory


mechanism. Particularly for stimuli at large distances, its sensitivity is superior for detecting movement in depth compared with change in disparity,
since the change in size is proportional to target size.
Change in disparity depends on the interocular distance, which is comparatively small for large target
distances. Depending on eye-target distance, different magnitudes of lateral and fore-aft sway are necessary to cause a net retinal displacement above the
threshold of motion perception. There is a linear
correlation between eye-target distance for lateral
sway, but a steep, non-linear increase of necessary
fore-aft sway with increasing target distance, especially beyond the grasping space (Paulus et aL 1984).
Recordings of postural sway on a force-platform
(Bronstein and Buckwell1997) or while walking on a
treadmill (Bardy et aL 1996) have shown that motion
parallax is used to control postural sway. Control of
fore-aft sway is based on optical expansion and
motion parallax in natural environments.

Visual stabilisation in the dark


A scotopic visual scene with few illuminated areas or
single objects must still provide visual information
for postural controL Kapteyn et aL (1979) have
shown that the degree of optokinetic destabilisation
of posture (by means of a tilting room) is a function
of illumination level within the state of transition
from scotopic to clear photopic vision (from 0.015
up to 1.3 lux), with saturation at 2.3 lux. The functional significance of this finding is that vision
affects posture, even under minimal illumination
conditions.
A single stationary target, provided by a lightemitting diode at a distance of 50 cm in an otherwise
dark environment, causes areduction in lateral sway
activity of ab out one-third in comparison with the
eyes-closed condition. This again indicates the powerful contribution of the fovea in the correction of
lateral body sway, whereas fore-aft sway is only
minimally stabilised because of the higher threshold
for detection of anterior-posterior movement
(Paulus et aL 1984). A complex visual environment
containing several structures is likely to provide
more information on head and body movement in
space than one containing only a single visual object.

Flicker illumination
With flicker illumination of the visual scene, it is
possible to evaluate the differential effects of continuous versus discrete visual inputs, which correspond,

427

Visual vertigo: visual contra I of motion and balance

respectively, to detection of retinal slip and retinal


dislocation of the viewed targets. With flicker frequencies of 16 and 32 Hz, body sway does not differ
significantly from that with continuous room illumination (Fig. 29.14). A flicker frequency of 8 Hz, however, causes an increase in body sway of about 30%,
whereas all frequencies tested below 8 Hz (1,2,4 Hz)
cause a common increase of instability of about 60%
(Paulus et al. 1984; Brandt et al. 1985).
Thus, with purely sequential retinal displacements of the visual scene due to flicker illumination,
a partial but significant visual stabilisation is preserved which, surprisingly enough, does not depend
on flicker frequencies between 1 and 4 Hz. As soon
as continuous motion perception becomes involved
(> 4 Hz), visual stabilisation gradually improves
with increasing frequencies (Fig. 29.14).
The flicker data substantiate and extend earlier
investigations of optokinetic destabilisation (Kapteyn
et al. 1979). They imply physiologically that although
continuous motion of the retinal image may be an
important determinant of postural stabilisation, it is
not a necessary precondition for it. Visual stabilisation of body sway was tested in a patient with severe
deficits of the vestibular system (due to gentamicin
treatment) and the somatosensory system (due to
polyneuropathy) (Paulus et al. 1987). With eyes open,
the patient was able to stand and walk slowly; with
eyes closed, he lost his balance within one second.
Under conditions of flicker illumination with
decreasing frequencies, he needed at least a visual

RMS

[0/0]

input rate of 17 Hz in order to maintain an upright


body position (Fig. 30.3, p. 448).

Visual field
That it is possible to balance and to walk while reading demonstrates that focal and ambient visual functions can be dissociated. Even though pattern
recognition and attention are dominated by the
reading material, orientation in space and postural
control are carried out unconsciously and adequately
on the basis of peripheral visual stimuli (Leibowitz
et al. 1982). We hypothesised that the high visual
acuity of a small central field may be counterbalanced by the wide-angle motion sensitivity of the
parafoveal and more peripheral retinal areas, which
subserve self-motion perception (linearvection, circularvection) relative to the surroundings (Brandt et
al. 1973).
Visual stabilisation of posture, however, differs
from circularvection-inducing mechanisms in that
central areas of the visual field (as compared to the
peripheral retina) playa dominant role in "finetuning" postural control. Under normal everyday
conditions, the visual field provides enough redundancy to compensate for small scotomas. This
explains the possibility of almost 100% balance performance with reduced visual field loss, e.g. with
monocular vision or exclusion of the fovea within
the "Ganzfeld;' although the fovea, if stimulated

FLASH - LIGHT ILLUMINATION


FORE - AFT BODY SWAY

300

250 200 -

---------------------------------------------------------r- - .----- .---r---

150 100 - --- , - - - - -

---

1---

-r-

--- r-r-

16 Hz

32 Hz

50 -

EYES OPEN

1 Hz

2Hz

4 Hz

8 Hz

EYES CLOSED

Fig.29.14. Fore-aft body sway during stroboscopic illumination of the surroundings with varying frequencies fram 1 to 32 Hz. With
flicker frequencies from 1 to 4 Hz, a constant partial stabilisation is preserved, based on information about retinal displacement. As
soon as continuous motion perception becomes involved (> 4 Hz), visual stabilisation gradually improves with increasing frequencies,
based on retinal slip information. (From Brandt et al 1985.)

Vertigo

428

alone, may contribute significantly. Foveal stabilisation becomes particularly apparent for correction of
smaIllateral (not fore-aft) body motion relative to
structures within the nearby grasping space
(Straube et al. 1994; Nougier et al. 1997). Patients
with field defects (e.g. central scotomas in multiple
sclerosis, homonymous hemianopia, tunnel vision in
retinitis pigmentosa) have a considerably reduced
postural balance; this becomes evident in the neurological examination when the patient walks tandem
or balances on one foot. It has been reported that the
cortical representation of retinal areas decreases
with increasing distance from the fovea: cortical
magnification factor (Virsu et al. 1982). Central and
paracentral visual fields have different thresholds for
detecting motion, but show the same contribution to
postural balance with correction for the cortical
magnification factor (Straube et al. 1994).

Eye movements, oculomotor


disorders, and postural balance
The detection of head motion relative to the surroundings is the basic cue for visual stabilisation of
posture. It requires detection of either the retinal slip
of the visual scene caused by head motion (afferent
motion perception) or compensatory eye movements that maintain fixation of a target within the
stationary scene (efferent motion perception). As
simple as it is, this distinction between afferent and
efferent motion perception seems to cover the mechanism of visual stabilisation of posture; it does not,
however, refiect natural stimulation. Under normal
conditions, this simple model is complicated by continuous voluntary eye movements, such as saccades
between stationary targets or pursuit of a single
moving target within a structured environment. It
seems, therefore, reasonable to investigate postural
sway during voluntary eye movements, when visual
perception of body sway is worsened by either suppression of motion perception during saccades or
additional retinal slip of the visual scene due to
pursuit.
Normal subjects balance worse during saccades or
smooth pursuit. Eccentric horizontal fixation, with
gaze defiected 40 in either direction, increases sway
by about 30% (Brandt et al. 1986; Straube et al. 1989).
The corresponding figure for vertical gaze defiection
of 35 up or down is 20% (Fig. 29.15). When voluntary smaIl-amplitude horizontal saccades are made
at a frequency of 0.5 Hz (motion perception is suppressed during periods of eye movement), visual
stabilisation of posture is in general preserved. At

up35

RMS 150 (%)

let! 40

down 35

RMS 150 ("10)

fore-oft
o lateral
Fig.29.15. Visual stabilisation of posture with eccentric gaze.
Fore-aft and lateral body sway (root mean square values (%) )
with fixation of eccentric horizontal or vertical targets. The dotted circle represents the 100% RMS values with the gaze straight
ahead. When the gaze is 40 to the right or to the left and 35
upward or downward, postural sway increases by about 30% and
20%, respectively. Eccentric gaze also impairs motion perception
(From Brandt et al. 1986.)

larger amplitudes and high er frequencies, which


reduce the rate at which visual input can be used for
stabilisation, postural sway significantly increases
(Brandt et al. 1986; Straube et al. 1988). The differential effects of frequency and amplitude of saccadic
eye movements may partially explain the controversial data of other authors who describe - to our
understanding, a paradoxical - improvement of
sway stabilisation during saccadic eye movements
(Iwase et al. 1979; Kikukawa and Taguchi 1985; Oblak
et al. 1985).
Pursuit of a single moving target within a stationary visual environment significantly impairs postural
balance. This can be explained by the increased
threshold for motion detection with moving eyes
(Probst et al. 1984). Postural balance may be worse
than with eyes closed (Fig. 29.16) under conditions
of increasing angular velo city of pursuit or irregular
pursuit of target with varying velo city as weIl as
direction.

Nystagmus with oscillopsia impairs balance


Involuntary ocular oscillations, which override fixation such as periodic alternating nystagmus (Fig.
29.17), acquired pendular nystagmus, spasmus
nutans, superior oblique myokymia, or downbeat!
upbeat nystagmus, are a weIl-known cause of oscil-

Visual vertigo: visual control of motion and balance

kinetic, saccadic, and other normal eye movements;


this causes an illusory motion of the visual world to
be perceived, in particular if head movements occur.
Patients, for whom the device is useful when the
head is at rest, find it destabilising when they move
about, and this applies equally to normal subjects
when wearing the device (Rushton et al. 1989).

RMS
(%)

horizontalsaccades (0.5 Hz)

400

80

.,>
~ 300

429

::J

~ 200

Cl..

100

fixation

saccades

eyes
closed

RMS
(%)

triangular pursuit (0.5 Hz)

400

.,>

. 300
.... 200
~

""

::J

Cl..

100

fixation
fore-aft

pursuit

eyes
closed

lateral

Fig.29.16. Postural sway during rapid and slow eye movements. Fore-aft and lateral body sway during voluntary horizontal saccades with amplitudes from 5 to 80 (top) or horizontal
pursuit from 5 to 80 at 0.5 Hz (bottom), compared with stationary fixation (RMS = 100%) and eyes closed.ln particular, pursuit of
a single moving target within a stationary visual environment
significantly impairs postural balance. (From Brandt et al. 1986.)

lopsia (see Oscillopsia, p. 430), independently of


head motion. Oscillopsia, the illusory movement of a
viewed stationary scene, differs from perception of
real motion in that it produces a disturbing experience of spatial disorientation, which affects posture
and locomotion (Brandt 1984). Both the apparent
motion of the stationary external visual scene and
the impaired motion perception (Bchele et al. 1983)
may cause areduction of visual stabilisation of posture in acquired ocular oscillations. In downbeat
nystagmus, the patient's pathological postural sway
with the eyes open depends on the direction of gaze;
it increases with increasing nystagmus amplitude,
secondary to reduced visual stabilisation owing to
the nystagmus (see Downbeat nystagmus syndrome,
p. 199). An optical device has been described
(Rushton and Rushton 1984) which provides partial
retinal image stabilisation. It was intended for use by
patients with continuous oscillopsia caused by nystagmus, which overrides fixation. The device also
stabilises the image during vestibulo-ocular, opto-

Extraocular muscle paresis impairs locomotion


and balance
The sudden onset of an extraocular muscle paresis,
e.g. in patients suffering from ocular myasthenia,
may induce "ocular vertigo" during voluntary eye or
head movements. The symptoms are not confined to
perceptual illusions such as oscillopsia but also
include postural imbalance and disturbance of locomotion. The impairment of motor control can be
attributed to an acute disturbance of the ability to
use vision to locate objects in egocentric coordinates,
an ability that requires knowledge of the position of
the eye relative to the head as well as the retinallocation of the image. Visually guided motor performance requires accurate information about gaze
direction as well as body position relative to the surroundings. Paretic deviation of normal eye position
due to the lack of afferent extra-retinal information
causes a dissociation of the subjective visual and
somatosensory"straight ahead", resulting in a mismatch responsible for the direction-specific dis tortion of locomotion and reaching movements (Fig.
29.18) .
The patient depicted in Fig. 29.19 suffered an
acute weakness, confined mainly to the right medial
and inferior rectus muscles, which caused abduction
and upward deviation of the right eye. Reaching
movements of the hand when guided by the affected
eye were displaced downward and to the left, indicative of a distortion of the visual straight-ahead
upward and to the right. When attempting to walk
forward, the patient experienced an apparent downward tilt of the floor, dizziness, and adeviation of
locomotion to the left, which led to an irresistible fall
to the right. Posturographic measurements of lateral
and fore-aft body sway revealed increased sway
amplitudes during monocular vision with the affected
eye compared with the unaffected eye. Postural
imbalance was particularly apparent when voluntary
sinusoidal head movements were performed about
the vertical z-axis during free upright stance. As
oscillopsia is the result of a functional deficiency of
the vestibular ocular reflex, the corresponding
increase in body sway can be interpreted as the
visuospinal consequence of mismatch between
the actual retinal slip du ring head movements and

430

Vertigo
periodic alternating nystagmus

............
EOG

EOG

0.7 0.6 c

'E

--

-S
co

0.5 0.4 -

c.

0.3 -

~
Ul

0.2-

>
co

R-L

~ ~ A-P

~ ~ -=-=3

EOG

EOG

~ ~ A-P
.0

R- L

0.1 0.0 -

gaze stra ight


ahead

lateral gaze
no oscillopsia

lateral gaze
osci llopsia

'-

'-----------~v~-----------/
eyes open

eyes closed

Fig.29.17. Head and body sway path in relation to nystagmus amplitude in a patient suffering from periodic alternating nystagmus.
Original horizontal electronystagmographic recordings are depicted for gaze straight-ahead and lateral gaze of 40 to the right and to
the left (top); the columns represent the sway path of body and head (bottom); postural sway increases with increasing nystagmus and
oscillopsia; all depend on the direction of gaze. (From Straube et al. 1989.) EOG = electro-oculogram

(J --'~

visual
- straight ahead

/1':.-.... .
<!--"\~."
.)\
..

above, ocular motility does not improve. This is due


to central nervous system plasticity and is consistent
with the reports of sensory and motor adaptation to
prolonged optical reversal of vision by prisms
(Kohler 1956; Melvill-Jones 1977).

Oscillopsia

Fig.29.18. Dissociation of subjective visual and somatosensory straight-ahead (monocular) vision in a patient suffering from
an acute extraocular muscle paresis with deviation of normal eye
position due to myasthenia gravis. (From Brandt and Bchele
1979.)

the expected pattern calibrated prior to the onset of


the disease. Postural imbalance with acute extraocular muscle paresis is particularly apparent when
voluntary head movements are performed or during
intentional gaze in the direction of the optimal range
of action of the paretic muscle (Fig. 29.19).
Although head exercises promote sensory
rearrangement causing subsequent disappearance of
the perceptual and postural symptoms described

Patients with external eye muscle paresis are often


unable to recognise faces or to read while walking,
and report motion of stationary visual scenes during
head motion or locomotion. This phenomenon is
termed oscillopsia (Brickner 1936). Oscillopsia also
occurs in diseases that cause involuntary ocular
oscillations (acquired pendular nystagmus, downbeat nystagmus, superior oblique myokymia)
(Bender 1965). Either the deficiency of compensatory
eye movements (due to an inappropriate vestibuloocular reflex) or the deficiency of visual fixation
(due to ocular oscillation) causes undesired retinal
image motion with disturbing oscillopsia.
Angular amplitude of the apparent motion perceived by the subject, however, does not quantitatively
match the net retinal slip (Brandt 1982; Wist et al.
1983; Bchele et al. 1983). The dissociation between
the two can be explained by the combination of two

Visual vertigo: visual control of motion and balance

431

tressing oscillopsia. The disadvantageous side effect


is in general impaired motion perception.
right eye fixating

feh eye occluded

bodV swav

loreaft

ga:r:e slfaight ahead

20Nm~

CD

20Nm

gaze upward right

Iree stance w ith head in normal posi t ion

Fig.29.19. Posturography of the fore-aft (block columns) and


lateral (stippled columns) body sway during free upright stance in
a 45-year-old male patient suffering from an acquired paresis of
the rectus superior muscle and an overaction of the obliquus
inferior muscle, indicated on the Hess-Lees screen (top). RMS values (N m) of body sway are minimal in the primary position of
gaze (1) and increase significantly when the gaze is directed 45
toward the optimal range of action of the affected extraocular
muscles (3,4). Original recordings of the fore-aft sway under gaze
conditions 1 and 4 are depicted at the bottom. (From Brandt
1984.)

separate mechanisms (Fig. 29.20), both of which


involve motion perception (Brandt and Dieterich
1988):
a physiological elevation of the threshold for
detecting object-motion with moving eyes and
2. a pathological elevation of the threshold for
detecting object-motion caused either by an
infranuclear ocular motor palsy or by supranuclear ocular oscillations
1.

In a teleological sense, this "adaptive suppression" of


the detection of retinal image motion is beneficial to
the organism to the extent that it aHeviates the dis-

Oscillopsia is smaller than retinal image slip:


deficient vestibulo-ocular reflex
In infranuclear eye-movement dis orders the perceived but illusory motion of the visual environment
is a consequence of inappropriate vestibulo-ocular
reflex (VOR) gain, which causes retinal image slip.
The VOR normaHy serves to hold constant the direction of gaze in space during head movements, by
driving the eyes to move in their orbits in the direction opposite to that of head motion, with a velo city
and amplitude that "compensate" for the head
motion. If the amplitude and/or velocity of eye
movements are inappropriate, the result is a shift in
the direction of gaze, causing a displacement or slip
of the retinal image, which may be perceived as an
apparent motion of the fixated object. As appealing
and simple as this model is, it is not fuHy supported
by experimental studies. It has been shown for
healthy subjects even under optimal fixation conditions - either with a biteboard, while sitting or
standing as still as possible (Stavenski et al. 1979), or
with head oscillations (Steinman and CoHewijn
1980) - that appreciable displacements of retinal
images result. Velocities of retinal slip ranged from
20 min of arc/s with the head fixed by a biteboard to
an average of 4/s with head oscillations. Moreover,
retinal image slip was found to be different in each
eye, with considerable relative binocular motion of
the fixated (objectively) stationary visual scene.
Despite such retinal image velocities, oscillopsia was
not reported. Thus, a stable world may be perceived,
even if the "compensatory" eye movements initiated
by the VOR are insufficient to cancel the retinal
image motion.
Bender and co-workers described patients in
whom oscillopsia was accompanied by, and presumably arose from, a deficient VOR, but they also
described cases in which oscillopsia existed in the
absence of any measurable oculomotor disturbances
(Bender 1965; Bender and Feldman 1967; Atkin and
Bender 1968; Gresty et al. 1977). Functional tests of
inappropriate compensatory eye movements have
been reported by Benson and Barnes (1978) using
visual acuity, by Wist et al. (1983) using a bedside
test of oscillopsia, as weH as by Zee (1978), who
observed relative movements of the optic disc by
means of ophthalmoscopy during active head movements.
To compare quantitatively the relationship
between oscillopsia and retinal slip, we recorded
head and eye movements simultaneously during

432

Vertigo

TWO MECHANISMS OF
IMPAIRED MOTION PERCEPTION
~

."

r-."

-~
!.~

~~

6r0

>
r-

l2

paretic eyes
cong. I acq. nystagmus

moving eyes
self - motion

DISORDERS
DEFICIENT VOR

I DEFICI~NT FIXAnON~

head movements

OSClllOPSIA

<

RETINAl SLIP

ocular oscillations

Fig.29.20. Oscillopsia is caused by either inappropriate compensatory eye movements (VOR) during head motion or oscillations
which override fixation. Amplitudes of perceived oscillopsia are always smaller than net retinal slip because motion perception under
these conditions is partially suppressed. Suppression of motion perception is due to the summation of physiological and pathological
(adaptive) elevation of thresholds to detect retinal image shifts of the viewed visual scene. (From Brandt and Dieterich 1988.)

sinusoidal oscillations about the vertical z-axis


( 20;1 Hz) while subjects fixated a stationary target
at a distance of 120 cm. For the VOR to be completely
compensatory (maintaining a constant gaze direction
in space), the eye must rotate with an angle exceeding that of the head, since the eye and head have different axes (Fig. 29.21). The additional angle of eye
motion () depends on the circumference of the
head (g) as weIl as the distance between the head
and the fixated target (a). In patients with subacute
peripheral ocular muscle paresis, the psychophysically determined angle of subjective oscillopsia is
considerably smaller than the calculated net retinal
slip or even absent altogether (Brandt 1982; Wist et
al. 1983). The patients are not able to detect the
amount of actual displacement of the visual scene on
the retina.

Acquired ocular oscillations with oscillopsia


Oscillopsia mayaIso occur in dis orders involving
involuntary ocular movements which override fixation in the absence of concomitant head motion, and
therefore do not involve the VOR. Here excessive

rather than deficient eye movements lead to the retinal


motion of images of stationary objects. The retinal
slip in downbeat nystagmus is misinterpreted as
motion of the visual scene, because the involuntary
ocular movements are not associated with an appropriate efference-copy signal.
Simultaneous psychophysical and electronystagmographic measurements have been performed in
patients with downbeat nystagmus to elucidate the
relationship between retinal image slip and oscillopsia (Bchele et al. 1983). Oscillopsia is a permanent
symptom, but the illusory motion is smaller than
would be expected from the amplitude of the nystagmus. Oscillopsia is dependent on the direction of
gaze, as is the amplitude of nystagmus, and increases
with increasing nystagmus amplitude, with a mean
ratio between the two of 0.37. The individual ratio is
relatively consistent for each patient, and interindividual values range from 0.13 to 0.61.
Thresholds for detection of egocentric objectmotion are significantly raised in patients with
downbeat nystagmus (Fig. 29.22; Brandt and
Dieterich 1988; Dieterich et al. 1998) in comparison
with healthy subjects; thresholds increase with
increasing nystagmus amplitude. Thus, there is a

433

Visual vertigo: visual control of motion and balance

c([deo]

50

"
"2

So

stationary target (Tl

20'

right
O

20'
20'

left
right

20:
10

left

O'

10'

40J\l\j\
43~

", : head oscillation


'2 :

eye oscillation

3 '~

1s

L---J

Fig. 29.21. Schematic representation of the experimental procedure for the measurement of eye-head coordination and retinal slip
during sinusoidal oscillatory head movements about the vertical z-axis with fixation of a stationary target. To maintain fixation on the
target during head oscillation, additional rotation of the eye is required (P), which results from the different axis of rotation of eye and
head. (u, angle of head rotation; u + p, angle of eye rotation; p, additional angle of eye rotation; g, head circumference (g, = 0.51 m, g2
= 0.61 m); u, target-z-axis distance (a = 1.2 m). Original recordings of head and eye movements when summated (S3) show that p is
equal to 3 in order to maintain gaze in space. (From Wist et al. 1983.)

partial suppression of perception of visual motion,


which affects both the retinal slip due to the involuntaryeye movements and single objects moving within the visual scene.

Physiological impairment of motion perception


with moving eyes
Perception of retinal image motion is suppressed by
normal (physiological) mechanisms during eye
movements. This contributes to the partial suppression of oscillopsia under perceptual conditions with
inappropriate compensatory eye movements or
acquired ocular oscillation. We have shown that the
thresholds for perception of egocentric objectmotion are significantly raised during concurrent
head oscillations and fixation of a target (Brandt
1982). Active sinusoidal head oscillations raised the
detection threshold for object-motion (24 min of
arc/s) by a factor of 2.9 at 1 Hz and 6.4 at 2 Hz (Fig.
29.23) despite intended stabilisation of the target on
the retina. Wertheim (1981) was able to demonstrate
that during smooth pursuit of a target (head stationary), the threshold for detection of motion of a visual

background increases in proportion to ocular velo city,


irrespective of whether the stimulus and the eyes
move in the same or opposite directions.
The phenomenon of suppressed visual motion
perception during eye movements may reflect a
basic sensorimotor mechanism because it has a
somatosensory analogue (Brandt and Dieterich
1988). Elevated thresholds for the perception of electrical stimuli applied to a fingertip as weIl as partially suppressed somatosensory evoked potentials are
reported during simultaneous movement of the
stimulated finger in humans (Coquery 1978;
Rushton et al. 1981). In the cat, the response in the
mediallemniscus to contralateral stimulation is suppressed, if the stimulus is applied less than 100 ms
before the onset of active movement (Ghez and Pisa
1972; Coulter 1973). This seems to support the
hypothesis of an efferent inhibition of sensory
inflow. Since suppression also occurs with passive
movements of the limbs, "afferent inhibition" must
also be possible (Angel and Malenka 1982).
Furthermore, analysis of EMG responses evoked in
the leg in humans by perturbations revealed that
monosynaptic stretch reflex responses as weIl as
supraspinal pathways of group I afferents are

434

Vertigo

100

congenital nystagmus (n = 5)
o downbeat nystagmus (n

=7)

.e

'
c

'E
50

vi

<J

"0

oe
CI)
cD
"-

...oe
0
Direction of gaze (eccentricity, deg)
Fig.29.22. Thresholds for detection of object-motion (24 min of arc/ s; mean s S.o.) as a function of the eccentricity of horizontal
gaze (0-40) in patients suffering from congenital nystagmus (n = 5) and acquired down beat nystagmus (n = 7) compared with normal
subjects (n = 12). Normal subjects show only a slight increase in thresholds with eccentric gaze, wh ich becomes more pronounced on
lateral gaze of 40. The thresholds for the patients are significantly raised, irrespective of whether the ocular oscillation is congenital or
acquired. There is a disproportional increase during lateral gaze beyond 20, which simultaneously activates nystagmus amplitude in
both disorders. (From Brandt and Oieterich 1988.)

min of are

200

180

i ....

~~

o-

.2.E

c
E~

" E
~~ 3

&

..

l!
~

160

I
I
r+l
head fixed

i1-

in

140 <l
120
-c-

100
80

,,!

"'t"

i}

1 Hz
2 Hz
head oscillation (t 20 )

60
40

iE

.
8

Ci
CI)
'6
Ci

e
c

20
0

Fig.29.23. Thresholds for detection of object-motion (means SO) in 12 normal subjects (columns) compared with eight patients
suffering from acquired peripheral ocular motor palsies (closed eircles, paretic eye; open eircles, unaffected eye). Ouring the measurements the subject fi xated the moving target (24 minutes of arc per second; left or right) while the head was either immobilised bya
biteboard or voluntarily oscillated about the vertical z-axis at 1 or 2 Hz with an amplitude of 20 (motion perception during vestibulo-ocular reflex). Physiologically, thresholds for detecting the motion of objects significantly increased in normal subjects with increasing frequency of head oscillation (columns). Patients with peripheral oculomotor palsies exhibited a further pathological elevation of
thresholds for both eyes under both the head-immobilised and head oscillation conditions.lt is obviously more pronounced in the
affected eye. (From Brandt and Oieterich 1988.)

435

Visual vertigo: visual control of motion and balance

suppressed during gait (Dietz 1986). Chapin and


Woodward (1981) have speculated about an
inhibitory interaction within the cortex itself
(between motor area 4 and somatosensory area 3),
since the reduction of afferent signals was more pronounced in cortex neurons than in corresponding
nuclei within the spinal cord.

Normal (physiological) inhibitory interactions


between self-motion and object-motion
perception
The chance observation of considerable difficulty
seeing treetops moving in the wind while driving a
vehicle led us to a systematic study of egocentric
perception of the motion of objects during selfmotion. In aseries of laboratory experiments we
found significantly increased thresholds for objectmotion perception during simultaneous self-motion
under conditions involving vestibular, optokinetic,
or cervicoproprioceptive stimulation (Brandt 1982;
Probst et al. 1984,1986). The results resemble the
phenomenon described in the previous section.
Wertheim (1994) has presented a theory that
describes the interfaces between self-motion and
object-motion precepts.
Real motion of the eyes or the head is not the
essential stimulus for suppression of object-motion
perception. This was demonstrated by slow oscillations of the trunk relative to the head (cervical stimulation), which was fixed by a biteboard (Probst et al.
1984,1986), and by circularvection studies (Brandt
1982) with objectively stationary subjects. When
apparent self-motion was visually induced in the
latter by full-field optokinetic stimulation with the
head fixed, the subjects experienced horizontal
optokinetically induced circularvection (see
Circularvection, p. 409).
The normal physiological inhibitory interaction
between object-motion and self-motion perception
may reftect a lack of specificity (or a side effect) of a
space constancy mechanism (efference-copy?),
which provides us with a stable picture of the world
during locomotion. It has practical implications
when one is riding in a vehicle and has the twofold
perceptual task of controlling self-motion (preferably by linearvection) and simultaneously perceiving object-motion. Critical changes in intervehicle
distances can take at least 300 ms longer to detect
than laboratory data suggest (obtained with the subject's head immobilised with a biteboard) (Probst et
al. 1984). Authorities on road traffic accidents should
take note! This hypothesis was proven in a field
study (vehicle guidance under natural conditions)
and with corresponding stimulation in the laboratory,

in which stationary surroundings eliminated the


perception of self-motion. These corrected detection
times thus call for an alteration of our concept of
safe intervehicle distance in a convoy.

Pathological (adaptive?) binocular impairment


of motion perception caused by monocular
external eye muscle paresis
The amount of net retinal slip tolerated by patients
(Fig. 29.23) without causing oscillopsia and the
interindividual differences with respect to the degree
and the acuteness of the paresis led us to suspect an
additional pathological impairment of motion perception. The latter is dependent on the particular
disease and is unrelated to the physiological
phenomena described above (Brandt and Dieterich
1986; Dieterich and Brandt 1987).
The patients had significantly raised thresholds
for the detection of object-motion with the head
fixed - by up to a factor of 5 for the paretic eye and a
factor of 3.3 for the normal eye. During sinusoidal
head oscillations at 1 Hz (which physiologically
elevates threshold) the ratio between threshold in
patients and normals was still about 4 for the
paretic eye and 2.7 for the unaffected eye (Fig. 29.23;
Brandt and Dieterich 1986). This clearly suggests
that both the physiological and the pathological
impairment of motion perception in patients with
acquired peripheral ocular motor palsies summate
when the moving target is fixated during voluntary
head motion. A central mechanism that affects
motion perception must be assumed, since perception with both eyes is involved, although the paretic
eye tends to perform more poorly. Impairment of
motion perception (its underlying mechanism is not
known) lasts as long as the palsy. The amount of
suppression of motion perception seems, for the
most part, to be independent of the degree of the
individual palsy, but the time course shows improvement of motion perception during the recovery of
the palsy.
A direction-specific impairment of vertical
motion perception was also described in downgaze
palsy due to rostral midbrain infarction (Heide et al.
1990).

Oscillopsia and motion perception in


congenital nystagmus
Oscillopsia is widely suppressed in congenital nystagmus (CN). It is usually absent in the primary
position of gaze and the null-zone of nystagmus, but
is still detectable by most sufferers as a subtle

Vertigo

436

oscillation of fixa ted objects when eccentric gaze


precipitates maximal amplitudes of the nystagmus.
This raises the question of possible impairment of
motion perception in CN, since in acquired downbeat nystagmus partial suppression of oscillopsia
was linked to impaired motion perception in general
(Fig. 29.22). It was the typical defect of optokinetic
nystagmus in CN which stimulated Kommerell and
Mehdorn (1982) to speculate that a dysfunction of
the feedback system controlling shifts of the images
across the retina might be the basic abnormality
underlying the development of CN. From their
experiments on estimating the magnitude of perceived motion, in which the patients managed to differentiate velocities within the range of 15 to 100 0 /s
almost as precisely as normal controls, they infer
that dysfunction of the feedback system is not simply caused by a sensory defect (Kommerell et al.
1986). However, because of the method used, their
data do not support the conclusion that motion perception is normal in these individuals. The method
(magnitude estimation) implies that if the standard
stimuli were perceived by the CN patients to be slower (than perceived by normals), it is still possible that
discrimination within the velo city range of stimulation is preserved. Thus, the power function of velocity perception would appear normal despite the
pathology of the CN patients, who underestimated
absolute velocity of all single pattern motions.
On the basis of our own data on the threshold of
motion detection in CN, we have reason to believe
that there is in fact a general impairment of motion
perception in these individuals (Dieterich and
Brandt 1987). Further studies will be required to
determine whether this threshold increment results
only in underestimation of velocity, or whether the
ability to predict and track the position of a moving
target is also impaired. Oscillopsia, and consequently
the sensitivity of the ability to detect retinal image
slip due to the nystagmus, is obviously less in CN
than in downbeat nystagmus, which suggests more
powerful "adaptation" in the congenital abnormality.
If the CN patients were able to subtract their current
eye motion from the change in position of a viewed
target on the retina (by efference-copy mechanism)
one would expect them to see an after image oscillate
in darkness in accordance with their nystagmus.
Surprisingly enough, this was neither observed by
von Hofe (1941) nor in particular by Godde-Jolley
and Larmande (1973), who reported on patients
with CN who could fixate a stationary object with a
foveal afterimage without seeing either the subject
or the after image oscillate. In contrast, Kommerell et
al. (1986) found some patients with CN who
observed oscillation of an afterimage in darkness
with an amplitude about half of the nystagmus

amplitude. The latter fits observations by Leigh et al.


(1988), in which oscillopsia (completely suppressed
under natural conditions) was precipitated in some
individuals with CN by using a combined lens device
that produces an approximate stabilisation of vision
of the real world despite movement of the eyes. This
device consists of two parts: a field lens and a
strongly divergent contact lens. The field lens is
placed so that its own principal focus is at the centre
of rotation of the eyeball and a ray from the fixation
point traverses the optical axis of the eye, regardless
of the direction in which the eye is pointing
(Rushton and Rushton 1984). The powerful suppression of oscillopsia in chronic ocular oscillation, as
weH as the recurrence of oscillopsia with a stabilised
retinal image (which interrupts continuous motion
stimulation), is reminiscent of the well-known velocity habituation and motion aftereffects following
prolonged motion stimulation (Brandt et al. 1974a).
CN affects perception of object-motion much more
severely than perception of self-motion; this indicates that adaptation to CN occurs on the level of
visual-vestibular interactions for the perception of
visual object-motion and not on the level of visual
motion signals (Eggert et al. 1997). Temporally
directed deficits for the detection of visual motion
were described in latent nystagmus (ShalloHoffmann et al. 1996).
In conclusion, undesired retinal image slips of the
fixated visual scene are caused by a deficient VOR
during head movements and by involuntary ocular
oscillations that override fixation when the head is
stationary. In all instances of ocular motor disorders
investigated so far (either supranuclear or infranuclear, either congenital or acquired), the amplitude
of perceived motion of the visual scene is considerably smaller than the calculated net retinal slip. This
partial suppression of distressing oscillopsia is
inevitably linked to impaired motion perception in
general (Fig. 29.20). Anormal physiological mechanism impairs motion perception with moving eyes
and contributes to the suppression of oscillopsia but
does not completely account for it. An additional
adaptive binocular impairment of motion perception must be involved, which is separate from the
physiological phenomenon and initiated by the particular oculomotor disorder.

References
Adler FH (1941) Ocular vertigo. Am Acad Ophthalmol
OtolaryngoI46:27-32
Angel RW, Malenka Re (1982) Velocity dependent suppression of
cutaneous sensitivity during movement. Exp Neurol 77:266-274

Visual vertigo: visual control of motion and balance


Arslan M, Molinari GA (1965) Modifications of the activity of the
vestibular nuclei in the cat following stimulation of the temporallobe. Acta Otolaryngol (Stockh) 59:338-344
Atkin A, Bender MB (1968) Ocular stabilization during oscillatory
head movements. Arch Neurol (Chicago) 19:559-566
Aubert H (1886) Die Bewegungsempfindung. Arch Ges Physiol
39:347-370
Bardy BG, Warren WH, Kay BA (1996) Motion parallax is used to
control postural swayduringwalking.Exp Brain Res 111:27l-282
Barret GV, Thornton CL (1968) Relationship between perceptual
style and simulator sickness. J Appl Psychol 52:304-308
Baumeyer F (1953) Der Hhenschwinde!. Nervenarzt 24:467-474
Bender MB (1965) Osciliopsia.Arch Neurol (Chicago) 13:204-213
Bender MB, Feldman M (1967) Visual illusions during head movement in lesions of the brain stern. Arch Neurol (Chicago)
17:354-364
Benson AI, Barnes GR (1978) Vision during angular oscillation:
the dynamic inter action of visual and vestibular mechanisms.
Aviat Space Environ Med 49:340-345
Berthoz A, Pavard B, Young LR (1975) Perception of linear horizontal self-motion induced by peripheral vision (linearvection). Basic characteristics and visual-vestibular interactions.
Exp Brain Res 23:471-489
Bles W, Kapteyn TS, DeWit G (1977) Effects of visual-vestibular
interaction on human posture. Adv Oto-Rhino-Laryngol
22:111-118
Bles W, Kapteyn TS. Brandt Th, Arnold F (1980) The mechanism
of physiological height vertigo. I!. Posturography. Acta
Otolaryngol (Stockh) 89:534-540
Bottini G, Sterzi R, Paulesu E, Vallar G, Cappa SF, Erminio F,
Passingham RE, Frith CD, Frackowiak RSJ (1994) Identification
of the central vestibular projections in man: a positron emission tomography activation study. Exp Brain Res 99:164-169
Brandt Th (1982) The relationship between retinal image slip,
oscillopsia and postural imbalance. In: Lennerstrand G, Zee DS,
Keller EL (eds) Functional basis of ocular motility disorders.
Pergamon Press, Oxford, pp 379-385
Brandt Th (1984) Visual vertigo and acrophobia. In: Dix MR,
Hood JD (eds). Vertigo. Wiley, Chichester, pp 439-466
Brandt Th, Bchele W (1979) Ocular myasthenia: visual disturban ce of posture and gait. Agressologie 20:195-196
Brandt Th, Daroff RB (1980) The multisensory physiological and
pathological vertigo syndromes. Ann NeuroI7:195-203
Brandt Th, Dieterich M (1986) Peripheral ocular motor palsy
impairs motion perception. In: Keller EL, Zee DS (eds) Adaptive
processes in visual and oculomotor systems. Pergamon Press,
Oxford, pp 457-463
Brandt Th, Dieterich M (1988) Oscillopsia and motion perception.
In: Kennard C, Clifford Rose F (eds) Physiological aspects of
dinical neuro-ophthalmology. Chapman & Hall, London, pp
321-339
Brandt Th, Wist ER, Dichgans J (1971) Optisch induzierte PseudoCoriolis-Effekte und Circularvektion: Ein Beitrag zur optischvestibulren Interaktion. Arch Psychiat Nervenkr 214:365-389
Brandt Th, Dichgans J, Knig E (1973) Differential effects of central versus peripheral vision on egocentric and exocentric
motion perception. Exp Brain Res 16:476-491
Brandt Th, Dichgans J, Bchele W (1974a) Motion habituation:
inverted self-motion perception and optokinetic after-nystagmus. Exp Brain Res 21:337-352
Brandt Th, Dichgans J, Wagner W (197 4b) Drug effectiveness on
experimental optokinetic and vestibular motion sickness.
Aerospace Med 45:1291-1297
Brandt Th,Arnold F, Bles W, Kapteyn TS (1980) The mechanism of
physiological height vertigo. I Theoretical approach and psychophysics. Acta Otolaryngol (Stockh) 89:513-523
Brandt Th, Esser I, Bchele W, Krafczyk S (1982) "Visuo-spinal
ataxia" caused by disorders of eye movements. In: Roucoux A,

437
Crommelinck M (eds) Physiological and pathological aspects of
eye movements. W Junk, The Hague, pp 425-430
Brandt Th, Paulus W, Straube A (1985) Visual acuity, visual field
and visual scene characteristics affect postural balance. In:
Igarashi M, Black FO (eds) Vestibular and visual control of posture and locomotor equilibrium. Karger, Basel, pp 93-98
Brandt Th, Paulus W, Straube A (1986) Vision and posture. In: Bles
W, Brandt Th (eds) Disorders in posture and gait. Elsevier,
Amsterdam,pp 157-176
Brandt Th, Bartenstein P, Danek A, Dieterich M (1998) Reeiprocal
inhibitory visual-vestibular interaction: visual motion stimulation deactivates the parieto-insular vestibular cortex. Brain
121:1749-1758
Brickner R (1936) Oscillopsia: a new symptom commonly occurring in multiple sderosis. Arch Neurol Psychiatr (Chicago)
36:586-589
Bronstein AM (1995) Visual vertigo syndrome: dinical and posturography findings. J Neurol Neurosurg Psychiatry 59:472-476
Bronstein AM (1996) Visually induced paroxysmal nausea and
vomiting as presenting manifestation of multiple sclerosis. J
Neurol Neurosurg Psychiatry 60:701
Bronstein AM, Buckwell D (1997) Automatie control of postural
sway by visual motion parallax. Exp Brain Res 113:243-248
Bchele W, Brandt Th, Degner D (1983) Ataxia and oseillopsia in
downbeat-nystagmus vertigo syndrome. Adv Oto-RhinoLaryngoI30:291-297
Bttner U, Buettner UW (1978) Parietal cortex (2v) neuronal
activity in the alert monkey during natural vestibular and optokinetic stimulation. Brain Res 153:392-397
Bttner U, Henn V (1981) Circularvection: psychophysics and single unit recordings in the monkey. Ann NY Acad Sei
374:274-283
Chapin JK, Woodward DJ (1981) Modulation of sensory responsiveness of single somatosensory cortical cells during movement and arousal behaviours. Exp Neurol 72: 164-178
Cheung BSK, Howard IP, Nedzelski JM, Landolt JP (1989)
Circularvection about earth-horizontal axes in bilateral
labyrinthine-defective subjects. Acta Otolaryngol (Stockh)
108:336-344
Cheung BSK, Howard IP, Money KE (1991) Visually-induced sickness in normal and bilaterally labyrithine-defective subjects.
Aviat Space Environ Med 62:527-531
Coquery J-M (1978) Role of active movement in control of afferent input from skin in cat and man. In: Gordon G (ed) Active
touch: the mechanism of recognition of objects by manipulation. Elmsford, New York, pp 161-169
Coulter JD (1973) Sensory transmission through lemniscal pathway during voluntary movement in the cat. J Neurophysiol
37:831-845
Crampton GH, Young FA (1953) The differential effects of a rotary
visual field on susceptibles and nonsusceptibles to motion sickness. J Comp Physiol Psychol46:45 1-453
Darwin E (1794) Zoonomia or, the laws of organic life. Voll, of
Vertigo. J. Johnson, London, pp 227-239
Davidson PW, Whitson TT (1973) Some effects of texture density
on visual cliff behaviour of the domestic chick. J Comp Physiol
Psychol 84:522-526
De Hardt DC (1969) Visual cliffbehaviour ofrats as a function of
pattern size. Psychonom Sci 15:268-269
Desnoes PH (1926) Seasickness. JAMA 86:319-324
Dichgans I, Brandt Th (1973) Optokinetic motion sickness and
pseudo-Coriolis-effects induced by moving visual stimuli. Acta
Otolaryngol (Stockh) 76:339-348
Dichgans J, Brandt Th (1978) Visual-vestibular interaction: effects
on self-motion perception and postural contro!. In: Held R,
Leibowitz HW, Teuber H-L (eds) Handbook of sensory physiology, vol8. Perception. Springer, Berlin Heidelberg New York, pp
755-804

438
Dichgans J, Schmidt CL, GrafW (1973) Visual input improves the
speedometer function of the vestibular nuclei in the goldfish.
Exp Brain Res 18:319-322
Dichgans J, Diener HC, Brandt Th (1974) Optokinetic-graviceptive interaction in different head positions. Acta Otolaryngol
(Stockh) 78:391-398
Dichgans ), Held R, Young LR, Brandt Th (1972) Moving visual
scenes influence the apparent direction of gravity. Seience
178:1217-1219
Dichgans ), Mauritz KH, Allum )HJ, Brandt Th (1976) Postural
sway in normals and atactic patients: analysis of the stabilizing
and destabilizing effects of vision. Agressologie 17:15-24
Diener HC, Wist ER, Dichgans J, Brandt Th (1976) The spatial frequency effect on perceived velocity. Vision Res 16: 169-176
Dieterich M, Brandt Th (1987) Impaired motion perception in
congenital nystagmus and acquired ocular motor palsy. Clin
Vision Sei 1:337-345
Dieterich M, Grnbauer WM, Brandt T (1998) Direction-specific
impairment of motion perception and spatial orientation in
downbeat and upbeat nystagmus. Neurosci Let! 245:29-32
Dietz V (1986) Afferent and efferent control of posture and gait.
In: Bles W, Brandt Th (eds) Disorders of posture and gait,
Elsevier, Amsterdam, pp 69-81
DiZio P, Li W, Lackner JR, Matin L (1997) Combined influences of
gravitoinertial force level and visual field pitch on visually perceived eye level.) Vestib Res 7:381-392
Droulez J, Berthoz A, Vidal pp (1985) Use and limits of visual
vestibular interaction in the control of posture. In: Igarashi M,
Black FO (eds) Vestibular and visual control on posture and
locomotor equilibrium. Karger, Basel, pp 14-21
Dupont P, Orban GA, De Bruyn B, Verbruggen A, Mortelmans L
(1994) Many areas in the human brain respond to visual
motion. J Neurophysiol72:1420-1424
Edwards AS (1946) Body sway and vision ) Exp Psychol
36:526-535
Eggert T, Straube A, Schroeder K (1997) Visually induced motion
perception and visual control of postural sway in congenital
nystagmus. Behav Brain Res 88:161-168
Fischer MH, Kornmller EE (1930) Optokinetisch ausgelste
Bewegungswahrnehmungen und optokinetischer Nystagmus. )
Psychol NeuroI41:273-308
Friberg L, Olsen TS, Roland PE, Paulson OB, Lassen NA (1985)
Focal increase of blood flow in the cerebral cortex of man during vestibular stimulation. Brain 108:609-623
Gantchev GN, Draganova N, Dunev S (1985) Influence of the
stabilogram and statokinesigram visual feedback upon the
body oscillations. In: Igarashi M, Black FO (eds) Vestibular and
visual control on posture and locomotor equilibrium. Karger,
Basel, pp 135-138
Ghez G, Pisa M (1972) Inhibition of afferent transmission in
cuneate nucleus during voluntary movement in the cat. Brain
Res 40:145-151
Gildenberg PL, Hassler R (1971) Influence of stimulation of the
cerebral cortex on vestibular nuclei units in the cat. Exp Brain
Res 14:77-94
Godde-Jolly D, Larmande A (1973) Les Nystagmus. Masson, Paris
Gramowski K-H (1962) Die Bedeutung der Vestibularisprfung
bei Tauglichkeitsuntersuchungen von Hhenarbeitern. HNO
10:279-281
Graybiel A (1970) Susceptibility to acute motion sickness in blind
persons. Aerospace Med 41 :650-653
Gresty MA, Hess K, Leech J (1977) Disorders of the vestibuloocular reflex producing oscillopsia and mechanisms compensating for loss of labyrinthine function. Brain 100:693-716
Groen JJ (1961) The problems of the spinning top applied to the
semicircular canals. Confin Neurol (Basel) 21:454-455
Grsser OJ, Pause M, Schreiter U (1982) Neuronal responses in the
parieto-insular vestibular cortex of alert Java monkeys (Macaca
jascicularis). In: Roucoux A, Crommelinck M (eds)

Vertigo
Physiological aspects of eye movements. W. Junk, The Hague,
pp 251-270
GFsser 0), Pause M, Schreiter U (1990a) Localization and
responses of neurons in the parieto- insular vestibular cortex of
the awake monkeys (Macaca jascicularis). ) Physiol
430:537-557
Grsser OJ, Pause M, Schreiter U (1990b) Vestibular neurons in
the parieto-insular cortex of monkeys (Macaca jascicularis):
visual and neck receptor responses. ) Physiol 430:559-583
Heide W, Fahle M , Koenig E, Dichgans J, Schroth G (1990)
Impairment of vertical motion detection and downgaze palsy
due to rostral midbrain infarction. Neurology 237:432-440
Helmholtz H von (1896) Handbuch der physiologischen Optik.
Voss, Leipzig
Henn V, Young LR (1975) Ernst Mach on the vestibular organ 100
years ago.) Otorhinolaryngol37: 138-148
Hofe K von (1941) Untersuchungen ber das Verhalten eines zentralen optischen Nachbildes bei und nach unwillkrlichen
Bewegungen sowie mechanischen Verlagerungen des Auges.
Graefes Arch OphthalmoI144:164-169
Hu S, Glaser KM, Hoffman TS, Stanton TM, Gruber MB (1996)
Motion sickness susceptibility to optokinetic rotation correlates to past history of motion sickness. Aviat Space Environ
Med 67:320-324
Hu S, Davis MS, Klose AH, Zabinsky EM, Meux SP, )acobsen HA,
Westfall )M, Gruber MB (1997) Effects of spatial frequency of a
vertically striped rotating drum on vection-induced motion
sickness.Aviat Space Environ Med 68:306-311
Iwase Y, Uchida T, Hashimoto M, Takegami T, Suzuki N (1979)
Body sway stabilization induced during saccadic eye movement. Postural Refl Body Equilibr 1:123-129
James W (1882) The sense of dizziness in deaf-mutes. Am) Otol
4:239-254
Kapteyn TS, Bles W (1977) Circularvection and human posture.
Relation between the reactions to various stimuli. Agressologie
18:335-339
Kapteyn TS, Bles W, Brandt Th, Wist ER(1979) Visual stabilization
of posture: effect of light intensity and stroboscopic surround
illumination. Agressologie 20: 191-192
Kennedy RS, Stanney KM (1996) Postural instability induced by
virtual reality exposure: development of a certification protocol. Int J Hum Comp Interact 8:25-47
Kennedy RS, Hettinger L), Harm DL, Ordy )M, Dunlap WP (1996)
Psychophysical scaling of circular vection (CV) produced by
optokinetic (0 KN) motion: individual differences and effects of
practice. J Vestib Res 6:331-341
Khan OA, Sandoz GM, Olek MJ (1995) Visually induced paroxysmal
nausea and vomiting as presenting manifestations of multiple
sclerosis (letter). J Neurol Neurosurg Psychiatry 59:342-343
Kikukawa M, Taguchi K (1985) Characteristics ofbody sway during saccadic eye movement in patients with peripheral vestibular disorders. In: Igarashi M, Black FO (eds) Vestibular and
visual control on posture and locomotor equilibrium. Karger,
Basel, pp 335-359
Kobrak F (1924) ber den Bergschwindel und andere praktisch
wichtige Schwindelphnomene. Mschr Ohrenheilk 58: 126-134
Kohler I (1956) Die Methode des Brillenversuches in der
Wahrnehmungspsychologie mit Bemerkungen zur Lehre der
Adaptation. Z Exp Angew PsychoI3:381-417
KommereIl G, Mehdorn E (1982) Is an optokinetic defect cause of
congenital nystagmus? In: Lennerstrand G, Zee DS, Keller EL
(eds) Functional basis of ocular motility dis orders. Pergamon
Press, Oxford, pp 159-167
KommereIl G, Horn R, Bach M (1986) Motion perception in congenital nystagmus. In: Keller EL,Zee DS (eds) Adaptive processes in visual and oculomotor systems. Pergamon Press, Oxford,
pp 485-491
Lee DN,Aronson E (1974) Visual proprioceptive control of standing in human infants. Perception Psychophys 15:529-532

Visual vertigo: visual control of motion and balance


Leibowitz HW (1955) The relation between the rate threshold for
the perception of movement and luminance for various durations of exposure. J Exp Psychol 49:209-214
Leibowitz HW, Shupert-Rodemer C, Dichgans J (1979) The independence of dynamic spatial orientation from luminance and
refractive error. Perception Psychophys 25:75-79
Leibowitz HW, Post RB, Brandt Th, Dichgans J (1982) Implications
of recent developments in dynamic spatial orientation and
visual resolution for vehicle guidance In: Wertheim AH,
Wagenaar WA, Leibowitz HW (eds) Tutorials on motion perception. Plenum Press, New York, pp 231-260
Leigh R1, Dell'Osso LF, Yaniglos SS, Thurston SE (1988)
Oscillopsia, retinal image stabilization and congenital nystagmus. Invest Ophthalmol Vis Sei 29:279-282
Lestienne F, Soechting JF, Berthoz A (1977) Postural readjustments induced by linear motion of visual scenes. Exp Brain Res
28:363-384
Lishman JR, Lee DN (1973) The autonomy of visual kinaesthesis.
Perception 2:287-294
Mach E (1875) Grundlinien der Lehre von den
Bewegungsempfindungen. Engelmann, Leipzig
Melvill-Jones G (1977) Plasticity in the adult vestibulo-ocular
reflex arc. Phil Trans R Soc Lond 278:319-334
Miller JW, Goodson JE (1960) Motion sickness in a helicopter simulator. Aerospace Med 31:204-212
Money KE (1970) Motion sickness. Physiol Rev 50:1-39
Mueller C, Kornilova L, Wiest G, Deecke L (1994) Visually induced
vertical self-motion sensation is alte red in microgravity adaptation. J Vestib Res 4:161-167
Nashner LM (1985) Strategies for organization ofhuman posture.
In: Igarashi M, Black FO (eds) Vestibular and visual control of
posture and locomotor equilibrium. Karger, Basel, pp 1-8
Nougier V, Bard C, Fleury M, Teasdale N (1997) Contribution of
central and peripheral vision to the regulation of stance. Gait
Posture 5:34-41
Oblak B, Gregoric M, Gyergyek L (1985) Effects of voluntary eye
saccades on body sway. In: Igarashi M, Black FO (eds)
Vestibular and visual control on posture and locomotor equilibrium. Karger, Basel, pp 122-126
Paige GD (1994) Senescence of human visual-vestibular interactions: smooth pursuit, optokinetic, and vestibular control of eye
movements with aging. Exp Brain Res 98:355-372
Parker DM (1971) A psychophysiological test for motion sickness
susceptibility. J Genet Psychol 85:87-92
Paulus W, Straube A, Brandt Th (1984) Visual stabilization of posture: physiological stimulus characteristics and clinical aspects.
Brain 107:1143-ll63
Paulus W, Straube A, Brandt Th (1987) Visual postural performance after loss of somatosensory and vestibular function. J
Neurol Neurosurg Psychiatry 50:1542-1545
Probst Th, Brandt Th, Degner D (1986) Object-motion detection
by concurrent self-motion perception: psychophysics of a new
phenomenon. Behav Brain Res 22: l-ll
Probst Th, Krafczyk S, Brandt Th, Wist ER (1984) Interaction
between perceived self-motion and object-motion impairs
vehicle guidance. Science 225:536-538
Purkinje JE (1820) Beitrge zur nheren Kenntnis des Schwindels
aus heautognostischen Daten. Med JB (sterreich) 6:79-125
Reason JT, Diaz E (1971) Simulator sickness in passive observers.
Ministry of Defence, FPRC report no. 1310. HMSO, London
Redfern MS, Furman JM (1994) Postural sway of patients with
vestibular dis orders during optic flow. J Vestib Res 4:221-230
Regan D, Beverly Kl (1979) Binocular and monocular stimuli for
motion in depth: changing disparity and changing size feed the
same motion-in-depth stage. Vision Res 19:1331-1392
Rennert H (1990) Hhenschwindel, Hhenangst und
Hhenphobie. Psychiat Neurol med PsychoI42:333-339
Romberg MH (1846) Lehrbuch der Nervenkrankheiten des
Menschen. Duncker, Berlin

439
Rushton DN, Rushton RH (1984) An optical method for
approximate stabilization of vision of the real world. J Physiol
(Lond) 357:3P
Rushton DN, Rothwall IC, Craggs MD (1981) Gating of
somatosensory evoked potentials during different kinds of
movements in man. Brain 104:465-491
Rushton DN, Brandt Th, Paulus W, Krafczyk S (1989) Postural
sway during retinal image stabilization. J Neurol Neurosurg
Psychiatry 52:376-381
Schubert G (1931) ber die physiologischen Auswirkungen der
Corioliskrfte bei Trudelbewegungen des Flugzeuges. Acta
Otolaryngol (Stockh) 16:39-47
Shallo-Hoffmann J, Faldon ME, Acheson JF, Gresty MA (1996)
Temporally directed defieits for the detection of visual motion
in latent nystagmus: evidence for adaptive processing. Neuroophthalmology 16:343-349
Stavenski AA, Hansen RN, Steinman RH, Winterson BJ (1979)
Quality of retinal image stabilization during small natural and
artifieial body rotations in man. Vision Res 19:675-653
Steinman RM, Collewijn H (1980) Binocular retinal image motion
during active head rotation. Vision Res 20:415-429
Straube A, Brandt Th (1987) Importance of the visual and vestibular cortex for self-motion perception in man (circularvection).
Human NeurobioI6:2ll-218
Straube A, Brandt Th, Probst Th (1987) Importance of the visual
cortex for postural stabilization: inferences from pigeon and
frog data. Human NeurobioI6:39-43
Straube A, Paulus W, Quintern 1, Brandt Th (1988) Visual ataxia
induced by eye movements: posturographic measurements in
normals and patients with ocular motor disorders. Clin Vision
Sei 4:107-113
Straube A, Btzel K, Hawken M, Paulus W, Brandt Th (1989)
Postural control in the elderly: differential effects of visual,
vestibular and somatosensory input. In: Amblard B, Berthoz A,
Clarac F (eds) Posture and gait: development, adaptation and
modulation. Elsevier, Amsterdam, pp 105-ll4
Straube A, Paulus W, Brandt T (1990) Influence ofvisual blur on
object-motion detection, self-motion detection and postural
balance. Behav Brain Res 40: 1-6
Straube A, Krafcyzk S, Paulus W, Brandt T (1994) Dependence of
visual stabilization of postural sway on the cortical magnification factor of restricted visual fields. Exp Brain Res 99:501-506
Travis RC (1945) An experimental analysis of dynamic equilibrium. J Exp PsychoI35:216-234
Ungs TJ (1989) The occurrence of the vection illusion among helicopter pilots while flying over water. Aviat Space Environ Med
60: 1099-11 0 1
Virsu V, Rovamo I, Laurinen P, Nsnen R (1982) Temporal contrast sensitivity and cortical magnification. Vision Res
22:1211-1217
Walk RD, Gibson EG (1961) A comparative and analytical study of
visual depth perception. Psychol Monogr 75 (15, whole no. 519)
Walk RD, Walters CP (1974) Importance of texture-density preferences and motion parallax for visual depth discrimination by
rats and chicks. J Comp Physiol PsychoI86:309-315
Walk RD, Gibson EJ, Tighe TJ (1957) Behaviour oflight-and-darkreared rats on a visual cliff. Science 126:80-81
Wenzel R, Bartenstein P, Dieterich M, Danek A, Weindl A,
Minoshima S, Ziegler S, Schwaiger M, Brandt Th (1996)
Deactivation of human visual cortex during involuntary ocular
oscillations: a PET activation study. Brain 119:101-ll0
Wertheim AH (1981) On the relativity of perceived motion. Acta
Psychol 48:97 -110
Wertheim AH (1994) Motion perception during self-motion: the
direct versus inferential controversy revisited. Behav Brain Sei
17:293-355
Wist ER, Brandt Th, Krafczyk S (1983) Oscillopsia and retinal slip:
evidence supporting a clinical test. Brain 106:153-168

440
Witkin HA ( 1949) Perception of body position and of the position of the visual field. Psychol Monogr 63: 1-46
Wood RW (1895) The "haunted swing" illusion. Psychol Rev
2:277-278
Young LR, Oman CM, Dichgans J (1975) Influence of head orien-

Vertigo
tation on visually induced pitch and roll sensation. Aviat Space
Environ Med 46:264-268
Zee DS (1978) Ophthalmoscopy in examination of patients with
vestibular disorders. Ann NeuroI3:373-374

Somatosensory vertigo

Somatosensory signals from musculotendinous


receptors in the neck and joints provide an accurate
kinesthetic feedback of the extent of head and limb
movements. These signals contribute to the perception of self-motion during active locomotion by converging with vestibular and visual input on
multimodal neurons in the vestibular nuclei and
thalamus, which project to cortical multisensory
areas in the parietal lobe, e.g. area 7 (see Chap. 13;
p. 219). Experimental studies in animals and
humans have confirmed the functional significance
of arthrokinetic input for arthrokinetic nystagmus
and self-motion sensation (p. 446). Questions relevant for the discussion of "somatosensory vertigo"
are whether and how the lack or inadequate release
of somatosensory input leads to vertigo or disequilibrium. Ataxia and unsteadiness co-occurring with
sensory polyneuropathy (p. 447) are readily recognised and gene rally explained by a deficient sense of
lower limb joint position. In contrast, dizziness and
unsteadiness suspected to be of cervical origin (socalled cervical vertigo, a controversial disorder of
questionable clinical significance which is diagnosed
too often) may be due to inadequate or unadapted
excess stimulation of neck receptors in cervical pain
syndromes.

Cervical vertigo
Cervical vertigo (CV) is unlike other vertigo syndromes' and because of this is surrounded by controversy. Neck afferents not only assist the
co ordination of eye, head, and body, but they also
affect spatial orientation and control of posture. This
implies that stimulation of, or lesions in, these structures can produce CV. In fact, unilaterallocal anaesthesia of the upper dorsal cervical roots induces
ataxia and nystagmus in animals, and ataxia without
nystagmus in humans. If CV exists outside these
441

experimental conditions, it is obviously characterised by ataxia and unsteadiness of gait, and not by
a clear rotational or linear vertigo. Neurological,
vestibular, and psychosomatic dis orders must first
be excluded before the dizziness and unsteadiness in
cervical pain syndromes can be attributed to a cervical origin. Here the use of dynamic posturography
and analysis of postural reflexes rather than
electronystagmography may prove helpful for hardening a diagnosis of suspected cv.
Todate, however, the syndrome remains only a
theoretical possibility awaiting a reliable clinical test
to demonstrate its independent existence (Brandt
1996). Vertigo can be accompanied by cervical pain,
and associated with head injury or whiplash injury;
and in some cases it improves dramatically with
physiotherapy. None of these instances provides convincing evidence of a cervical mechanism, and alternative explanations are possible (p. 445).
Supporters of cervical vertigo usually believe it to
be the most common vertigo syndrome; they confirm their diagnosis with a range of signs, symptoms, and tests (nystagmus induced by head
rotation, for example), which are either irrelevant or
inappropriate.
Their opponents reject the diagnosis for two
reasons. In the first place, there is neither a reliable
clinical test for the syndrome nor a typical time
course for the condition. Second, reliable and wellestablished signs and tests can support a convincing
alternative diagnosis in about 90% of patients presenting with vertigo (Table 30.1).
The ongoing fierce controversy between the
believers in CV and the non-believers tends to
obscure the evidence and the issues. The following
discussion will cover the functional significance of
neck reflexes and their effects on eye movements
and posture. In addition, the findings in human and
in animal studies on nystagmus and ataxia produced
by dysfunction of neck afferents will be examined,
and the current views on CV in the clinicalliterature
critically reviewed (Brandt 1996).

442
Table 30.1.

Vertigo
Cervical vertigo (experimental and clinical evidence)

Experimental CV (unilateral suboccipital anaesthesia)


Rabbit, cat, monkey
Ataxia with ipsilateral deviation of gait
Positional nystagmus
(Contraversive horizontal nystagmus)
Humans
Sensation of floating and unsteadiness of gait
Ataxia with increased ipsilateral and decreased contra lateral extensor
muscle tone
Ipsiversive tendency to fall and deviation of gait
Ipsiversive past-pointing
No spontaneous or positional nystagmus
Clinical CV
Clinical syndrome
Ataxia and unsteadiness associated with some neck pain and/or
limitation of neck movement
No pathognomonic or reliable clinical test available (dynamic
posturography may be helpful)
Incidence/age/sex
Unknown
Pathomechanism
Unadapted somatosensory tone imbalance of upper two cervical
root inputs resulting in multisensory mismatch of sensorimotor
control for balance
Management
Physical and medical treatment of cervical pain syndrome
Differential diagnosis
- Traumatie vertigo (central or peripheral vestibular; otolith;
perilymph fistula)
- Cerebellar or spinal ataxia
- Phobie postural vertigo
- Bilateral vestibulopathy
- Vestibular paroxysmia
- Benign paroxysmal positioning vertigo

Functional significance of neck afferents and


neck reflexes
Spatialorientation
It is necessary for the sensorimotor control system

to know the attitude of the head relative to the body,


since the vestibular system only signals head motion
relative to space. The head-mounted sensory systems
must transform the rotations and accelerations they
sense and correctly relate their direction to the
motion and attitude of the body and the centre of
gravity. Neck afferents provide information about
the head attitude, and make an important contribution to the control ofbody and sensory spatial orientation. The perception of head or trunk rotations in
space would be erroneous, if only vestibular stimulation or only neck stimulation was involved. However,
if the two stimuli are combined (head rotation relative to the trunk), the perception ofboth trunk and
head rotation in space reflects the true position
(Mergner et al. 1991). Trunk rotations relative to the

stationary head (neck proprioceptive stimulation)


evoke three different tuming sensations, depending
on which aspect of the stimulus or wh at part of the
body the subject focuses his attention on:
a sensation of trunk rotation in space in the
direction of, and proportional to, the actual
trunk rotation;
2. an illusory sensation of head movement in space
opposite to the direction of the trunk rotation;
and
3. asensation of head sensation relative to the
trunk (Mergner et al. 1983a,b).
1.

Vestibular and visual cues produce postural corrections. They change their direction in egocentric
spatial coordinates with changes in head position
(Brandt et al. 1981). When the head is rotated horizontally (yaw) by 90 to the right or left, for example,
horizontal head accelerations and horizontal retinal
slip of the visual scene (right-Ieft in head coordinates) no longer indicate lateral body sway; instead,
they represent fore-aft movements. Consequently,
the compensatory postural adjustments must be corrected by neck afferences to reflect the change in
coordinates. Theoretically, unilateral irritation or
lesional deficit of neck input could cause abilateral
tone imbalance, thus disturbing integration of visualvestibular stimulation (head) and neck input (body).
Thus it is reasonable to investigate the perceived
apparent straight ahead and the subjective vertical
in patients with suspected cv. Unilateral electrical
stimulation of the neck (Wapner et al. 1951) or trunk
tilt relative to the head, which was fixed in space
(Fischer 1927), cause deviation of the subjective vertical. Vibration of muscles stimulates the primary
endings of the muscle spin dies, increases their firing
rate (as if the muscle is being stretched), and thereby
elicits a tonic contraction of muscle (Matthews
1966), giving the kinaesthetic illusion of head or
limb movement. Unilateral vibration of the posterior
neck muscles elicits an apparent head motion
(Taylor and McCloskey 1991) and an apparent movement of a visual target (Biguer et al. 1989) to the contralateral side. Accordingly, subjective "straight
ahead" shifts toward the side of the posterior neck
muscle vibration in healthy subjects and patients
with hemi-neglect (Kamath 1994). The section of
upper cervical nerve roots or muscles or local anaesthesia (De Jong et al. 1977; Dieterich et al. 1993)
obviously cause a somatosensory cervical tone
imbalance with respect to CV. However, it has not
been convincingly demonstrated that whiplash
injuries or cervical pain syndromes produce such a
tone imbalance with ataxia and vertigo. The probable mechanism also remains obscure.

Somatosensory vertigo

Major movements of the head in different directions occur about two different joints in the cervical
column: head rotations about Cl/C2 and head flexion and extension about C6/C7. The segment
between these two joints is considerably more rigid,
allowing only limited flexion, extension, lateral tilt,
and rotation.
Proprioception is not a function of the superficial
neck muscles (Hinoki and Terayama 1966) but of the
deep short intervertebral neck muscles, which are
extensively supplied with muscle spin dIes (Voss
1958; Cooper and Daniel 1963; Richmond and
Bakker 1982). The significant role played by muscle
spindies in cervical reflexes was supported by findings from experimental injections of succinyl
choline (Mergner et a1. 1982; Chan et a1. 1987).
Despite the high density of spindies in deep neck
muscles, the precision of subjective evaluation of
head position does not exceed that of the positional
sense in the joints of the extremities (Taylor and
McCloskey 1988). The Pacini receptors of the periarthricular tissue are obviously of less importance,
since the joint-position sense is only slightly reduced
after total hip replacement (Grigg et a1. 1973). Golgi
tendon organs and skin receptors may add some
information to the sensing of position and movements of joints in humans (McCloskey 1978).
Neck reflexes

Two reflexes are mediated by neck proprioceptors:


the postural neck reflexes and the cervico-ocular
reflex. Tonic postural neck reflexes, described by
Magnus and De Kleyn (1912), Weiland (1912) and
Magnus (1924), innervate limb muscles asymmetrically. However, coactivation of the labyrinth and neck
reflexes results in reciprocal effects which enable
head movements to occur without causing postural
imbalance (von Holst and Mittelstaedt 1950; Roberts
1973; Lindsay et a1. 1976; Mergner et a1. 1983a,b). In
humans, tonic postural neck reflexes can be elicited
only in the newborn (Barimy 1918; Gesell 1938), for
example, ipsilateral flexion and contralateral extension of the limbs with head rotation ("fencing posture"). The gradual disappearance of these reflexes
through suppression (mediated by supramesencephalic structures) is considered a sign of maturity.
Pathological (unsuppressed) tonic postural neck
reflexes have been attributed to lesions of the frontomesencephalic pathways (Schaltenbrand 1925), e.g.
in anencephalic dysplasia ("mesencephalic monstra") (Gamper 1926). The neck not only modulates
body posture, but it also stabilises the head in space
by cervicocollic reflexes, which are similarly integrated with vestibulocollic reflexes (Peterson et a1.
1985; Dutia and Hunter 1985; Wilson et a1. 1995). In

443

healthy human beings, neck reflexes form apart of


the multisensory postural control mechanism
(Stenvers 1936; Fukuda 1961), thus making it impossible for the clinician to carry out a selective test of
neck function by simple postural manoeuvres. Tonic
postural neck reflexes depend on the input of the
upper two (or three?) cervical nerves (De Kleyn
1921). Their section abolishes these reflexes in the
decerebrate, labyrinthectomised cat (McCouch et a1.
1951).
Cervico-ocular reflex

It was Barimy (1906) who first demonstrated tonic

cervico-ocular reactions in rabbits, elicited by


motion of the trunk relative to the head, and these
reflexes were later studied more thoroughly by
Magnus (1924). In humans, this tonic oculomotor
neck reflex can, like postural reflexes, only be
observed in the newborn (Barany 1918) or in rare
patients suffering from gross brainstem lesions
(Stenvers 1924; De Kleyn and Stenvers 1941).
Bikeles and Ruttin (1915) were the first to report
nystagmus during head rotation in patients with
complete vestibular loss, which they ascribed to sensory input from neck joints. Frenzel (1928)
described the "Schlagfeldverlagerung" as a neck
reflex and Gttich (1940) observed cervical nystagmus due to trunk rotation in healthy humans with
normallabyrinthine function, which was later confirmed by others (Bos and Philipszoon 1963; Collard
et a1. 1967; Jongkees 1969). The gain of the cervicoocular reflex in humans is low (- 0.3) compared to
that of the vestibulo-ocular reflex (Barnes and
Forbat 1979; Holtmann 1988; Sawyer et a1. 1994). It is
increased in acquired vestibular loss (Chap. 8,
p. 136), thereby partially compensating (substituting) for the vestibular deficit in the monkey
(Dichgans et a1. 1974) and in humans (Zangemeister
and Stark 1983; Bles et a1. 1984; Bronstein and Hood
1986). In contrast to that of the VOR, the gain of the
cervico-ocular reflex is difficult to determine, since it
is maximal with velocities of trunk rotation below
SOls and progressively decreases with increasing
velocity of rotation (Holtmann 1988).
Attempts to define and classify a pathological cervical nystagmus for the diagnosis of CV (Moser
1974; Hlse 1983; Scherer 1985; Hamann 1985) cannot succeed, since cervical nystagmus also occurs in
healthy subjects (Norre and Stevens 1987; Holtmann
1988).
Central pathways

The central pathways involved in neck reflexes,


wh ich affect eye and body muscles, have been

Vertigo

444

extensively studied in animal experiments (for


review see De Jong and Bles 1986). Section of the
brainstem caudally to the medial vestibular nuclei
abolishes the tonic cervico-ocular reflex in the rabbit
(Lorente de N6 1926); unilateral section of the dorsal
column elicits nystagmus in the cat (De Jong et al.
1977). Hikosaka and Maeda (1973) found oligosynaptic connections between neck afferents and
abducens motor neurons. These pathways ascend
ipsilaterally in the spinal cord, cross to the inferior
olive and the synapse in the vestibular nuclei to
become integrated with the VOR. There is an intense
visual-vestibular-proprioceptive convergence at all
levels of the central vestibular system, which is necessary for multisensory interaction in spatial orientation and postural control. Tonic postural neck
reflexes could be conveyed by cervical interneurons,
wh ich receive cortical and proprioceptive input
(Erulkar et al. 1966), or the central cervical nucleus
(Abrahams et al. 1984; Hirai et al. 1984; Fitz- Ritson
1985), or C3/C4 propriospinal neurons as described
by Illert et al. (1981).

Ataxia and nystagmus in experimental cervical


vertigo
Transverse section of suboccipital muscles results in
ataxia in animals (Longet 1845; Bernard 1865). This
was later confirmed by unilateral surgical deafferentation of Cl-C3 in the squirrel monkey (Igarashi
et al. 1969,1972) and in the cat (Manzoni et al. 1979)
as well as by local suboccipital anaesthesia in rhesus
monkeys (De Jong et al. 1977).
Local anaesthesia of deep posterolateral neck tissue, performed by Barre (1926), Schubert (1950),
Hinoki and Kurosawa (1964), and De Jong et al.
(1977), in humans usually elicits a transiently
increased ipsilateral and decreased contralateral
extensor muscle tone with a tendency to fall, and a
deviation of gait and past-pointing toward the
injected side. This was confirmed in patients with
cervicogenic headache investigated before and after
bilateral therapeutic anaesthetic C2-blockades
(Dieterich et al. 1993); however, Dieterich and coworkers found no specific abnormality with static
posturography, determination of subjective visual
vertical, or routine electronystagmography. The
weak horizontal spontaneous nystagmus, directed
away from the injected side, which was observed by
Barre (1926), is not a typical feature of this experimental condition in humans (De Jong et al. 1977).
However, electrical stimulation of the spinal nerves
can elicit eye movements in animals (Suzuki and
Takemori 1971).
Biemond (1939,1940) was the first to report posi-

tional nystagmus due to upper cervical root section


in the rabbit. Nystagmus was also induced by local
anaesthesia of neck afferents (Schubert 1950). Cohen
(1961) showed that this positional nystagmus is
species-specific: most pronounced in rabbits, less in
the cat, and subtle in the rhesus monkey. Positional
nystagmus cannot be attributed to a disturbance of
the cervical sympathetic chain as suggested by Barre
(1926). Animal experiments by Biemond and De
Jong (1969) and De Jong et al. (1977) indicate that it
most probably represents a tone imbalance of upper
cervical roots.
Perceptual and postural illusions can be induced
by applying vibrations to muscle tendons (Goodwin
et al. 1972), especially the posterior neck muscles
(Taylor and McCloskey 1991; Karnath 1994).
Similarly, nystagmus and motion sickness (p. 485)
can be provoked by regional vibration of the head
(Lackner and Graybiel1974).
The possible characteristics of CV are summarised in Table 30.1.

(linical evidence for cervical vertigo?


The frequency and severity of CV should not be
overestimated. Rotational vertigo and nystagmus
associated with pain arising from the cervical spine
with tenderness and limitation of neck movement
should not be called CV. More likely symptoms of
CV would be a sensation of numbness or floating
with unsteadiness and slight ataxia of stance and
gait. This can be inferred from the experimental vertigo induced in htimans by unilateral suboccipital
local anaesthesia (De Jong et al. 1977).
Since somatosensory cervical input converges
with vestibular input in order to mediate multisensory control of orientation, gaze in space, and posture, the clinical syndrome of CV should
theoretically include
1. perceptual symptoms of disorientation,
2. ocular motor signs, and
3. postural imbalance.
Consequently, further clinical studies on CV should
focus on establishing reliable measures for it. These
could include
1. psychophysics of perceived vertical, subjective
straight ahead, subjective head to trunk position,
or oscillopsia
2. oculographic determinations of gaze in space
and nystagmus
3. posturographic measurements, cervicospinal
reflexes, and analysis of gait.

445

Somatosensory vertigo

Since assessment of all these measures under static


conditions has so far proven inconclusive, further
investigations should focus on dynamic studies utilis.ing .vari.ous s~matosensory, vestibular, and optokmetIc stImulatIOns, such as trunk rotation with the
head stationary or, even more promising, vibration
of the upper posterior neck museles.
If vestibular function is tested by vestibular stimuli and visual function by visual stimuli, then
somatosensory cervical function should be tested
with somatosensory stimulation.
Up to now the most sensitive data - which are
however, not specific enough to establish the diagnosi~
- have bee~ obtai~ed with p.osturography (Karlberg
1995). PatIents wIth chromc cervicobrachial pain
syndrome (not selected for complaints of vertigo)
had significantly poorer postural control based on
vibration-induced and galvanically induced body
sway than sex- and age-matched controls (Karlberg
et al. 1995). In a consecutive study Karlberg and cowor kers \ K~r lberg et al. 1996a) tried to separate
charactenstIc patterns of postural control in
patients. with d~zziness of suspected cervical origin
and patIents wlth recent vestibular neuritis. Their
"posturographic parameters were 'swiftness', corresponding to integrative control wh ich reflects the
tim~ ~equired for realignment to the initial resting
positIOn after a perturbation. 'Stiffness', which
des~ri?es the reaction of the body on proportional
deVIatIOn ~rom the ass~med resting position, by
an~logy wI~h the functIon of aspring; 'damping'
WhIC~ descnbes the control action dependent on the
velocIty of the body-sway induced by a perturbation
cOI?parable to the effect of a shock absorber. High
sWlftness values mean rapid response to disturbance
and that the subject quickly returns to the chosen
resting position after a perturbation; high stiffness
values m:a~ that the subject reacts strongly to a
small deVIatIOn of body position and high damping
values mean fewer o~cillations of lower velo city
around the chosen restmg body position after a perturbation" (Johansson et al. 1988; Karlberg et al.
1996a). Patients with suspected CV were characterised by significantly lower values for stiffness and
sig~ificantly h~gh:r values for damping than healthy
subJects and slgmficantly lower values for stiffness
than the vestibular neuritis patients (Karlberg et al.
1996a). Some other postural tests were proposed,
such as head extension (DeJong and Bles 1986) or
the most prone head position (lund et al. 1993).
Routine static posturography (lund et al. 1991
~ieterich et al.. 1993) or nystagmus induced by pas~
Slve head rotatIons (Moser 1974; Hlse 1983; Scherer
1985) cannot elearly differentiate between normal
and diseased subjects.

Hypothetical mechanisms
It is not known how traumatic, degenerative, inflam~atory, or rheumatic diseases affect neck sensory
mp~t. In such uncharted regions, various hypotheses

e.g. the hypothesis of CV following whiplash


mJury. Several not very convincing causes have been
suggested: the neuromuscular mechanism (Gray
1956), the neurovascular mechanism (Hyslop 1952;
Weeks and Travelli 1955), or mechanical vascular
obstruction of the vertebral artery (Compere 1968).
Longet (1845) very early made the incidential obser:ration th~t post -traumatic vertigo and ataxia
lmprove with the use of a neck collar (Hinoki et al.
1971). But head trauma and whiplash injury do not
affect only neck structures. The otoliths are obviously more vulnerable to accelerations; damage to them
causes otolith vertigo (p. 349), which is characterised
by a benign course similar to that of neck pain.
Furthermore, whiplash injuries frequently damage
also the brain (Ommaya et al. 1968; Torres and
Shapiro 1961), making the interpretation of abnormal vestibulo-ocular tests difficult (Rubin 1973
Toglia 1976).
'
A convincing mechanism of CV would have to be
based on altered unilateral or bilateral upper cervical somatosensory input secondary to neck tenderness and limitation of movement. Interstitial
inflammatory mediators have been accused of sensitising musele spin dies (Johansson and Sojka 1991),
and myofascial trigger points exhibit spontaneous
EMG activity, which is compatible with hyperactive
musele spindies (Hubbard and Berkoff 1993). If the
firing characteristics (symmetrical or asymmetrical)
of .the cervi.cal somatosensors change due to neck
pam, a multlsensory mismatch would be expected to
result in CV. Since humans are not adapted to abnormal cervical input, the resultant mismatch would be
maximal during active head movements (when
expected and actual reafferent input do not match;
see ~lso Ch~p. 32, p. 471) and in the initial phase of
cervlcal pam syn~romes, before the multisensory
system rearranges ltself to the unusual input.
Three experimental approaches may come elose
to approximating a model of CV: restrained cervical
mobility b~ rigid neck collars (Karlberg et al. 1991),
lo~al ~ervlcal anaesthesia (DeJong et al. 1977;
Dletench et al. 1993), and neck muscle vibration
(Taylor and McCloskey 1991; Karnath 1994; Strupp
et al. 1998).
Cervical vertigo - reality or fiction? All clinical
~tud~e.s to date on CV have two weak points: (1) the
mablhty to confirm the diagnosis and (2) the unexplained discrepancy between patients suffering from
sev~r~ neck p~in without vertigo and patients complammg of dlsabling vertigo with moderate neck
pain. If CV exists, appropriate management is the
~h~lve,

Vertigo

446

same as that for the cervical pain syndrome, and this


management should not be denied any patient. Since
such therapy is carried out in any case, the heated
debate on the relevance and mechanism of CV is
more of theoretical neurophysiological interest than
of practical import. There are reports available that
demonstrate improvement of postural performance
in patients with dizziness of suspected cervical
origin following surgery of cervical root compression (Persson et al. 1996) or physiotherapy (Karlberg
et al. 1996b).
Differential diagnosis

Differential diagnosis of vertigo associated with cervical pain syndromes depends on the aetiology of
cervical pain. If it is post-traumatic or follows cervical whiplash injuries, then post-traumatic otolith
vertigo (p. 349), or benign paroxysmal positioning
vertigo (p. 251), central vestibular dysfunction secondary to brainstem concussion (see postconcussion
syndrome, p. 347), and perilymph fistulas should be
considered (see also Chapters 22 and 23; Table 30.1).
In non-traumatic cases, phobic postural vertigo
(p. 469) has a similar symptomatology, but cerebellar or spinal ataxia, immunological disorders,
vestibular paroxysmia (p. 117), and bilateral vestibulopathy (p. 127) should first be ruled out before cervical origin is assumed.

Arthrokinetic nystagmus and selfmotion sensation


Somatosensory input from muscle spindIes and joint
receptors provide kinaesthetic information about
limb movements and subserve our perception of
self-motion during locomotion. It has been demonstrated experimentally in humans that movement of
the limbs can induce a compelling illusion of selfmotion and a purely somatosensory nystagmus,
even in the absence of concurrent vestibular or optokinetic stimulation:
1. Arthrokinetic nystagmus is evoked in stationary

subjects seated in total darkness inside a rotating


cylinder, when they track the rotation of the
cylinder by placing their hands on its inner wall
(Fig. 30.1; Brandt et al. 1977).
2. Apparent "stepping around" nystagmus is evoked
in stationary subjects on a small circular treadmill in darkness (Bles 1981). This was later confirmed in the monkey and studied
physiologically with respect to stabilisation of

gaze during circular locomotion in light and


dark (Solomon and Cohen 1992a,b).
In our experiments on healthy subjects with passive
arm rotation at the shoulder joint, lateral nystagmus
with the fast phase beating in the opposite direction
to the arm movement was consistently found. The
mean slow phase velo city increased with increasing
arm velo city and reached about 15/s, the mean
position of the eyes deviated toward the fast phase as
in optokinetic nystagmus, and the nystagmus continued after the cessation of stimulation (arthrokinetic after-nystagmus) (Brandt et al. 1977). A
sensation of linear self-motion can be induced in the
blindfolded, stationary, sitting subject, who keeps
contact with a linearly moving platform (acceleration 0.1 m/s 2) in the frontal parallel plane by means
of a hand-over-hand walking action (Bles et al.
1995). When discordant suprathreshold vestibular
information from the otoliths is added by moving
the subject laterally (acceleration 0.1 m/s 2 ) in the
same direction as the platform (acceleration of the
platform 0.2 m/s 2 ), so that the arthrokinetic stimulus
is also accelerated by 0.1 m/s 2 but in the opposite
direction, the arthrokinetic information was found
to prevail on the perceived direction of self-motion
(Bles et al. 1995).
The existence of asensation of arthrokinetic selfmotion and nystagmus is a strong argument in
favour of a functionally significant somatosensoryvestibular convergence within the central vestibular
system, at least for afferents carrying positional and
kinesthetic information from the muscle spindies
and the joints. This convergence finally leads to activation of the neuronal circuits of the paramedian
pontine tegmentum, producing nystagmus and
mediating the sensation of body movement. It can be
assumed that somatosensory afferents playa powerful role within the multisensory process of selfmotion perception. Optimal functioning requires the
continuous evaluation of the reafferent sensory consequences of self-generated body movements, and a
mutually interactive calibration of the three main
loops: visual, vestibular, and somatosensory. When
the gain in smooth ocular tracking was investigated
for visual, vestibular, and arthrokinetic cues, jointly
as weIl as individually, evidence was found for
arthrokinetic and vestibular enhancement of
smooth ocular tracking (de Graaf et al. 1994).
Patients with bilateralloss of labyrinthine function (Chap. 8; p. 127) exhibit characteristic abnormalities of somatosensory (arthrokinetic) nystagmus
(Bles et al. 1983, 1984):

drastic shortening of latencies,


rapid build-up of slow phase velo city,

447

Somatosensory vertigo

Circularvection
in darkness

deg I s

20
10
0

rig.

STIMULUS
passive arm rotation

ARTHRO..NETIC ClRCULARVECTION

Fig.30.1. Arthrokinetic nystagmus. Schematic drawing of the stimulus condition (Ieft) and an example of an original recording of
arthrokinetic self-motion perception (circularvection) and nystagmus (right) induced by passive rotation of the arm of a stationary subject inside a rotating drum. Passive rotation of the arm to the right induces nystagmus and asensation of self-motion to the left.
(Brandt et al. 1977.)

increased gain and absence of stimulation afternystagmus.

These alterations can, on the one hand, be interpreted


as substitution for the loss of vestibular function,
and, on the other, as an increase of the particular
sensorial weight (importance) attributed to the
somatosensory information that provides selfmotion cues. In these patients the rapid onset and
build-up of somatosensory nystagmus may result
from the absence of vestibular signals. When, however,
they are present, they generate a signal that conflicts
with (inhibits) the somatosensory response. The
reversible changes in gain (increase-decrease) of the
cervico-ocular reflex in a patient who recovered
from post-meningitic bilateral vestibular loss supports this interpretation (Bronstein et al. 1995).

Other forms of nystagmus induced by nonvestibular stimulation


Nystagmus, an alternating sequence of quick phases
(saccades) and slow phases (eye tracking), can be
induced by vestibular or optokinetic stimulation. It
is also evoked by other stimuli: torsion of the cervical vertebral column (Bos and Philipszoon 1963;
Mergner et al. 1983b) - the sensation ofrotation is

more compelling with neck somatosensory than


with vestibular stimulation (Bles and DeJong 1982);
a moving sound source (von Stein 1910; Dodge 1923;
Hennebert 1960; Ganz et al. 1969); imagining a moving visual image (Zikmund 1966); or by hypnotic
suggestion of seen motion (Brady and Levitt 1964).

Postural imbalance with sensory


polyneuropathy
The "vertigo" described by patients with sensory
polyneuropathy is more a kind of ataxia, postural
imbalance, and unsteadiness, which occurs under
"scotopic" visual conditions at night or with the eyes
closed (Brandt et al. 1986; Kotaka et al. 1986; Paulus
et al. 1987).
Contral of posture and gait in humans is based on
the activation of preprogrammed patterns of muscle
activity, which are initiated and modulated by multiloop control pathways. These pathways carry information from visual, vestibular, and somatosensory
sources ab out the postural effects of this muscle
activity. In a healthy human being, these sensory
inputs converge, providing redundant information.
Thus, loss of information from one sensory input is
not critical, for example, during eye closure.
Important as it is for the dem an ding balancing tasks

Vertigo

448

in sports, it rarely manifests as clinically significant


instability (Brandt 1988). If, however, the sensorial
weight or reliability of the other modalities is reduced,
for example, by attempting to balance on one foot
with eyes closed and the head extended (see Head
extension and vertigo, p. 297), then a physiological
sensory imbalance can be demonstrated (Fig. 30.2).
In the absence of sensory information from two
of these systems, postural control may be severely
impaired, for example, in a patient with sensory
polyneuropathy under restricted visual conditions
(darkness) or with head extension, a position in
which the otoliths are beyond their optimal working
range (Fig. 30.2).
The importance of visual stabilisation of body
sway was studied in a patient who had severe deficits

of the vestibular (due to gentamicin treatment) and


the somatosensory systems (due to polyneuropathy),
but who was able to stand and walk slowly with eyes
open (Paulus et al. 1987). However, with eyes closed,
he lost balance within one second. His balance
performance also deteriorated when visual acuity
was experimentally reduced below 0.3 or during
flicker illumination (below 17 Hz) of the surroundings (Fig. 30.3; see Chap. 29, Visual stabilisation of
posture, p. 423).
Patients with postural imbalance due to sensory
polyneuropathy are rarely referred for assessment of
vertigo or dizziness. If so, the most relevant differential diagnosis is bilateral vestibulopathy (p. 127).
Both conditions can be distinguished by the sensory
loss in the lower limbs and the lack of oscillopsia in
sensory polyneuropathy. Phobie postural vertigo
(p. 469) may be a further differential diagnosis.

eyes open
head upright
20 Nm I
20 Nm

head extension

H P.d'.48

tore-alt

~----------------,~~~~~--~~

RMS

(./01

~ __,a_te_ra_,______________~___________

iii
~

100

100
evas closed

&
207

145

eyes open
20 Nm '

258

20Nm~

,..
...

174

16 Hz

~~-10S-Fig.30.2. Postural imbalance with sensory polyneuropathy.


Differential effects of head extension and normal head position
upon fore-aft and lateral body sway (original recordings), with
the eyes open or closed. Normal subject (27-year-old male)
standing on a firm posture platform (top); patient (57-year-old
female) with sensory polyneuropathy standing on a firm platform (bottom). Postural imbalance is the most pronounced for the
patient during head extension with the eyes closed, when there
is a combined absence of stabilising visual input and accurate
positional sense from ankle joint and muscle tendon receptors.
(Brandt et al. 1986.)

Fig.30.3. Postural imbalance in combined bilateral vestibular


failure and sensory polyneuropathy.Original recordings offoreaft and lateral body sway in a patient (48-year-old male) with
almost complete 1055 of labyrinth function (due to gentamicin
treatment) and lower limb joint position sense (due to sensory
polyneuropathy).ln contrast to steady room illumination (top),
body sway increases with decreasing flicker frequency under
stroboscopic illumination. At 16 Hz, balance cannot be maintained even for seconds, as was the case with eyes closed (bottom).
(From Paulus et al. 1987.)

Somatosensory vertigo

References
Abrahams VC, Richmond FJR, Keane J (1984) Projections from C2
and C3 nerves supplying muscles and skin of the cat neck: a
study using transganglionic transport of horseradish peroxidase. J Comp NeuroI230:142-154
lund M, Larsson SE, Ledin T, dkvist L Mller C (1991) Dynamic
posturography in cervical vertigo. Acta Otolaryngol (Stockh)
SuppI481:601-602
lund M, Ledin T dkvist L, Larsson SE (1993) Dynamic posturography among patients with common neck disorders. J Vestib
Res 3:383-389
Barimy R (1906) Augenbewegungen, durch Thoraxbewegungen
ausgelst. Zentralbl PhysioI20:298-302
Barany R (1918) ber einige Augen- und Halsmuskelreflexe bei
Neugeborenen. Acta Otolaryngol (Stockh) 1:97-103
Barnes GR, Forbat LN (1979) Cervical and vestibular afferent control of oculomotor response in man. Acta Otolaryngol (Stockh)
88:79-87
Barre JA (1926) Sur une syndrome sympathique cervical
posterieur et sa cause frequente: l'arthrite cervicale. Rev Neurol
45:1246-1253
Bernard C (1865) translated by HC Greene (1961) An introduction
to the study of experimental medieine. Collier Books, New York
Biemond A (1939) On a new form of experimental position nystagmus in the rabbit and its clinical value. Proc Kon Ned Akad
Wet 42:370-375
Biemond A (1940) Further observations about the cervical form
of positional-nystagmus and its anatomical base. Proc Kon Ned
Akad Wet 43:901-906
Biemond A, De Jong JMBV (1969) On cervical nystagmus and
related dis orders. Brain 92:437-458
Biguer B, Donaldson IML, Hein A, Jeannerod M (1989) Neck
muscle vibration modifies the representation of visual motion
and direction in man. Brain 111:1405-1424
Bikeles F, Ruttin E (1915) ber die reflektorischen kompensatorischen Augenbewegungen bei beiderseitiger Ausschaltung
des N. vestibularis. Neurol ZbI34:807-810
Bles W (1981) Stepping around: eircularvection and Coriolis
effect. In: Long JB, Baddeley AD (eds) Attention and performance, vol IX. Lawrence Erlbaum, Hillsdale, NI, pp 47-61
Bles W, Klren Th, Bchele W, Brandt Th (1983) Somatosensory
nystagmus: physiological and clinical aspects. Adv Oto RhinoLaryngoI30:30-33
Bles W, De Jong JMBV (1982) Cervico-vestibular and visuovestibular interaction. Acta Otolaryngol (Stockh) 94:61-72
Bles W, De Jong JMBV, Rasmussens J (1984) Postural and oculomotor signs in labyrinthine defective subjects. Acta
Otolaryngol (Stockh) SuppI406:101-104
Bles W, Jelmorini M, Bekkering H, de Graaf B (1995) Arthrokinetic
information affects linear self-motion perception. J Vestib Res
5:109-116
Bos JH, Philipszoon AJ (1963) Some forms of nystagmus provoked
by stimuli other than accelerations. Pract Oto- Rhino-Laryngol
25:108-118
Brady JP, Levitt EE (1964) Nystagmus as a criterion of hypnotically
induced visual hallucinations. Seience 146:85-86
Brandt Th (1988) Sensory function and posture. In: Amblard B,
Berthoz A, Clarac F (eds) Posture and gait: development adaptation and modulation. Elsevier, Amsterdam, pp 127-136
Brandt Th (1996) Cervical vertigo - reality or fiction? Audiol
Neurootoll:187-196
Brandt Th, Bchele W, Arnold F (1977) Arthrokinetic nystagmus
and ego-motion sensation. Exp Brain Res 30:331-338
Brandt Th, Krafczyk S, Malsbenden J (1981) Postural imbalance
with head extension: improvement by training as a model for
ataxia therapy. Ann NY Acad Sei 374:636-649

449
Brandt Th, Bchele W, Krafczyk S (1986) Training effects on
experimental postural instability: a model for clinical ataxia
therapy. In: Bles W, Brandt Th (eds) Disorders of posture and
gait. Elsevier, Amsterdam, pp 353-365
Bronstein AM, Hood JD (1986) The cervico-ocular reflex in normal subjects and patients with absent vestibular function. Brain
Res 373:399-408
Bronstein AM, Morland AB, Ruddock KH, Gresty MA (1995)
Recovery from bilateral vestibular failure: implications for visual
and cervico-ocular function. Acta Otolaryngol (Stockh) Suppl
520:405-407
Chan YS, Kasper J, Wilson VJ (1987) Dynamics and directional
sensitivity of neck muscle spindie responses to head rotation. J
NeurophysioI57:1716-1729
Cohen LA (1961) Role of eye and neck proprioceptive mechanisms in body orientation and motor coordination. J
Neurophysiol24: I-lI
Collard M, Conraux C, Thiebaut MS, Thiebaut F (1967) Le nystagmus d'origine cervicale. Rev NeuroI117:677-688
Compere WE (1968) Electronystagmographic findings in patients
with "whiplash injuries". Laryngoscope 78:1226-1233
Co oper S, Daniel PM (1963) Muscle spindies in man: their morphology in the lumbricals and the deep muscles of the neck.
Brain 86:563-586
de Graaf B, Bos JE, Wich S, Bles W (1994) Arthrokinetic and
vestibular information enhance smooth ocular tracking during
linear (self-)motion. Exp Brain Res 101:147-152
De Jong PTVM, De Jong JMBV, Cohen B, Jongkees LBW (1977)
Ataxia and nystagmus induced by injection of local anesthetics
in the neck.Ann Neuroll:240-246
De Jong JMBV, Bles W (1986) Cervical dizziness and ataxia. In:
Bles W, Brandt Th (eds) Disorders of posture and gait. Elsevier,
Amsterdam, pp 185-206
De Kleyn A (1921) Tonische Labyrinth - und Halsreflexe auf die
Augen. Pflgers Arch Ges PhysioI186:82-97
De Kleyn A, Stenvers HW (1941) Tonic neck reflexes on the eye
muscles in man. Proc Kon Ned Akad Wet 44:385-396
Dichgans I, Bizzi E, Morasso P, Tagliasco V (1974) The role of
vestibular and neck afferents during eye-head coordination in
the monkey. Brain Res 71:225-232
Dieterich M, Pllmann W, Pfaffenrath V (1993) Cervicogenic
headache: electronystagmography, perception of verticality and
posturography in patients before and after C2 -blockade.
Cephalalgia 13:285-288
Dodge R (1923) Thresholds of rotation. J Exp PsychoI6:107-137
Dutia MB, Hunter MJ (1985) The sagittal vestibulocollic reflex and
its interaction with neck proprioceptive afferents in the decerebrate ca!. J Physiol 359:17-29
Erulkar SD, Sprague JM, Whitsel BL, Dogan S, Jannetta PJ (1966)
Organisation of the vestibular projection to the spinal cord of
the cat. J NeurophysioI19:626-644
Fischer MH (1927) Messende Untersuchungen ber die
Gegenrollung der Augen und die Lokalisation der scheinbaren
Vertikalen bei seitlicher Neigung. Albrecht v Graefes Arch
OphthalmoII18:633-680
Fitz-Ritson D (1985) The direct connections of the C2 dorsal root
ganglia in the Macaca irus monkey: relevance to the chiropractic profession. J Manipulative Physiol Ther 8:147-156
Frenzel H (1928) Rucknystagmus als Halsreflex und
Schlagfeldverlagerung des labyrinthren Drehnystagmus durch
Halsreflexe. Z Hals Nasen Ohrenheilk 21:177-187
Fukuda T (1961) Studies on human dynamic postures from the
viewpoint of postural reflexes. Acta Otolaryngol (Stockh) Suppl
161:1-52
Gamper E (1926) Bau und Leistungen eines menschlichen
Mittelhirnwesens (Arhinencephalie mit Encephalocele).
Zugleich ein Beitrag zur Teratologie und Fasersystematik Ir.
Klinischer Teil. Z Gesamte Neurol Psychiat 104:49-120
Ganz H, Fend I, Huth FW (1969) Versuche ber audiokinetische

450
Augenbewegungen. 1. Binaurale Reizung bei Normalhrigen. Z
Laryngol Rhinol 49:625-636
Gesell A (1938) The tonic neck reflex in human infant. J Pediatr
l3:455-464
Goodwin GM, McCloskey DI, Matthews PBC (1972) The contribution of muscle afferents to kinesthesia shown by vibration
induced illusions of movement and by the effects of paralysing
joint afferents. Brain 95:705-748
Gray LP (1956) Extra-Iabyrinthine vertigo due to cervical muscle
lesions. J Laryngol 70:352-361
Grigg P, Finerman GA, Riley LH (1973) Joint-position sense after
total hip replacement. J Bone Joint Surg Am 55:1016-1025
Gttich A (1940) ber den Antagonismus der Hals- und
Bogengangsreflexe bei der Bewegung des menschlichen Auges.
Arch Ohr Nasen Kehl Kopf Heil Kd 147:1-4
Hamann KF (1985) Kritische Anmerkung zum sogenannten
zervikogenen Schwinde!. Laryngol Rhinol OtoI64:156-157
Henneber! PE (1960) Nystagmus audiocinetique. Acta
Otolaryngol (Stockh) 51:412-415
Hikosaka 0, Maeda M (1973) Cervical effects on abducens
motoneurons and their interaction with vestibulo-ocular reflex.
Exp Brain Res 18:512-530
Hinoki M, Kurosawa R (1964) Studies on vertigo provoked by
neck and nape muscles. Notes on vertigo of cervical origin.
Some observations on vertiginous attacks caused by injection
of procaine solution into neck and nape muscles in man. Oto
Rhino Laryngol Clin (Kyoto) 57:10-20
Hinoki M, Terayama K (1966) Physiological role of neck muscles
in the occurrence of optic eye nystagmus. Acta Otolaryngol
(Stockh) 62:157-170
Hinoki M, Hine, Kada Y (1971) Neurological studies on vertigo
due to whiplash injury. Equilib Res Suppll:5-29
Hirai N, Hongo T, Sazaki S (1984) A physiological study of identification, axonal course and cerebellar projection of spinocerebellar tract cells in central cervical nucleus of the cat. Exp Brain
Res 55:272-285
Holtmann S (1988) Die Analyse zerviko-okulrer Reaktionen
unter quantifizierten Reizbedingungen. Habilitationsschrift,
LMUMnchen
Hubbard DR, Berkoff GM (1993) Myofascial trigger points show
spontaneous ne edle EMG activity. Spine 18:1803-1807
Hlse M (1983) Die zervikalen Gleichgewichtsstrungen.
Springer, Berlin Heidelberg New York
Hyslop G (1952) Intra-cranial circulatory complication of injuries
of the neck. Bull NY Acad Med 28:729-733
Igarashi M, Alford BR, Watanabe T, Maxian PM (1969) Role of
neck proprioceptors for the maintenance of dynamic bodily
equilibrium in the squirrel monkey. Laryngoscope
79:17l3-1727
Igarashi M, Miyata H, Alford BR, Wright WK (1972) Nystagmus
after experimental cervicallesions. Laryngoscope 82:1609-1621
Illert M, Jankowska A, Lundberg A, Odutola A (1981) Integration
in descending motor pathways controlling the forelimb in the
cat. 7. Effects from the reticular formation on C3-4 propriospinal neurones. Exp Brain Res 42:269-281
Johansson H, Sojka P (1991) Pathophysiological mechanisms
involved in the genesis and spread of muscular tension in occupational muscle pain and in chronic musculoskeletal pain syndromes: a hypothesis. Med Hypotheses 35:196-203
Johansson R, Magnusson M, kesson M (1988) Indentification of
human postural dynamics. IEEE Trans Biomed Eng 35:858-869
Jongkees LBWC (1969) Cervical vertigo. Laryngoscope
79:1473-1484
Karlberg M (1995) The neck and human balance: a clinical and
experimental approach to "cervical vertigo". Thesis, University
of Lund, Sweden
Karlberg M, Magnusson M, Johansson R (1991) Effects of
restrained cervical mobility on voluntary eye movements and
postural contro!. Acta Otolaryngol (Stockh) 111:664-670

Vertigo
Karlberg M, Persson L, Magnusson M (1995) Reduced postural
control in patients with chronic cervicobrachial pain syndrome. Gait & Posture 3:241-249
Karlberg M, Johansson R, Magnusson M, Fransson PA (1996a)
Dizziness of suspected cervical origin distinguished by posturographic assessment of human postural dynamics. J Vestib Res
6:37-47
Karlberg M, Magnusson M, Malmstrm EM, Melander A, Moritz U
(1996b) Postural and symptomatic improvement after physiotherapy in patients with dizziness of suspected cervical origin.
Arch Phys Med Rehabil 77:874-882
Karnath HO (1994) Subjective body orientation in neglect and the
interactive contribution of neck muscle proprioception and
vestibular stimulation. Brain 117:1001-1012
Kotaka S, Croll GA, Bles W (1986) Somatosensory ataxia. In: Bles
W, Brandt Th (eds) Disorders of posture and gait. Elsevier,
Amsterdam, pp 178-183
Lackner JR, Graybiel A (1974) Elicitation ofvestibular side effects
by regional vibration of the head. Aerospace Med 45: 1267 -1272
Lindsay KW, Roberts TDM, Rosenberg JR (1976) Asymmetric
tonic labyrinth reflexes and the interaction with neck reflexes
in the decerebrate cat. J PhysioI261:583-601
Longet FA (1845) Memoires sur les troubles qui surviennent dans
l'equilibration,la station et allocomotion des animaux apres la
section des parties molles de la nuque. Gaz Med Paris
l3:565-567
Lorente de No R (1926) Die Grundlagen der Labyrinthphysiologie. Scand Arch PhysioI49:251-311
Magnus R (1924) KrpersteIlung. Springer, Berlin
Magnus R, De Kleyn A (1912) Die Abhngigkeit des Tonus der
Extremittenmuskeln von der KopfsteIlung. Pflgers Arch Ges
PhysioI145:455-584
Manzoni D, Pompeiano 0, Stampacchia G (1979) Cervical control
of posture and movements. Brain Res 169:615-619
Matthews PBC (1966) The reflex excitation of the soleus muscle of
the decerebrate cat caused by vibration applied to its tendon. J
PhysioI184:450-472
McCloskey DJ (1978) Kinesthetic sensibility. Physiol Rev
58:763-820
McCouch GP, Deering ID, Ling TH (1951) Location of receptors
for tonic neck reflexes. J NeurophysioI14:191-195
Mergner T, Anastasopoulos D, Becker W (1982) Neuronal responses to horizontal neck deflection in the group X region of the
cat's medullary brainstem. Exp Brain Res 45:196-206
Mergner T, Deecke L, Becker W, Kornhuber HH (1983a)
Vestibular-proprioceptive interactions: Neurophysiology and
psychophysics. In: v Horn (ed) Fortschritte der Zoologie, Bd.
28. Multimodal convergences in sensory systems. Fischer,
Stuttgart
Mergner T, Nardi GL, Becker W, Deecke L (1983b) The role of the
canal-neck interaction for the perception of horizontal trunk
and head rotation. Exp Brain Res 49: 198-208
Mergner T, Siebold C, Schweigart G, Becker W (1991) Human perception of horizontal trunk and head rotation in space during
vestibular and neck stimulation. Exp Brain Res 85:389-404
Moser M (1974) Zervicalnystagmus und seine diagnostische
Bedeutung. HNO 22:350-355
Norre ME, Stevens A (1987) Cervical vertigo. Acta Oto Rhino
Laryngol Belg 41:436-452
Ommaya AK, Faas, F, Yarnell P (1968) Whiplash injury and brain
damage. JAMA 204:285-289
Paulus W, Straube A, Brandt Th (1987) Visual postural performance after loss of somatosensory and vestibular function. J
Neurol Neurosurg Psychiatry 50:1542-1545
Persson L, Karlberg M, Magnusson M (1996) Effects of different
treatment on postural performance in patients with cervical
root compression. J Vestib Res 6:439-453
Peterson BW, Goldberg J, Bilotto G, Fuller JH (1985) Cervicocollic

Somatosensory vertigo
reflex: Its dynamic properties and interaction with vestibular
reflexes. J NeurophysioI54:90-109
Richmond FJR, Bakker DA (1982) Anatomical organization and
sensory receptor content of soft tissues surrounding upper cervical vertebrae in the cat. J NeurophysioI48:49-61
Roberts TDM (1973) Reflex balance. Nature 244:158-185
Rubin W (1973) Whiplash with vestibular involvement. Arch
Otolaryngol 97:85-87
Sawyer RN, Thurston SE, Becker KR, Ackely CV, Siedman SH,
Leigh RJ (1994) The cervico-ocular reflex of normal human
subjects in response to transient and sinusoidal trunk rotations. J Vestib Res 4:245-249
Schaltenbrand G (1925)
Normale Bewegungsund
Lagereaktionen bei Kindern. Dtsch Z Nervenheilkd 87:23-59
Scherer H (1985) Halsbedingter Schwindel. Arch Oto Rhino
Laryngol Supplll:107-124
Schubert K (1950) Schwindel und Sympathikus. Arch Ohr Nasen
Kehlk Heilkd 156:489-499
Solomon D, Cohen B (1992a) Stabilization of gaze during circular
locomotion in light. 1. Compensatory head and eye nystagmus
in the running monkey. J NeurophysioI67:1146-1157
Solomon D, Cohen B (1992b) Stabilization of gaze during circular
locomotion in light. Ir. Contribution of velocity storage to compensatory eye and head nystagmus in the running monkey. J
NeurophysioI67:1158-1169
Stenvers HW (1924) ber die klinische Bedeutung der kompensatorischen Augenbewegungen bei Kopfdrehung. Z Ges Neurol
Psychiat 92:484-486
Stenvers HW (1936) Haltungs- und Sttzreflexe. In: Bumke D,
Foerster 0 (eds) Handbuch der Neurologie, vol V/3.Allgemeine
Neurologie, Springer, Berlin, pp 523-554
Strupp M, Arbusow V, Dieterich M, Sautier W, Brandt Th (1998)
Perceptual and oculomotor effects of neck muscle vibration in
vestibular neuritis. Ipsilateral somatosensory substitution of
vestibular function. Brain 121:677-685

451
Suzuki J, Takemori S (1971) Eye movements induced from the
spinal nerves. Equilib Res SuppI2:33-40
Taylor JL, McCloskey DI (1988) Proprioception in the neck. Exp
Brain Res 70:351-360
Taylor JL, McCloskey DI (1991) Illusions of head and visual target
displacement induced by vibration of neck muscles. Brain
114:755-759
Toglia JV (1976) Acute flexion-extension injury of the neck.
Neurology 26:808-814
Torres F, Shapiro SK (1961) Electroencephalograms in whiplash
injuries.Arch NeuroI5:28-35
von Holst E, Mittelstaedt H (1950) Das Reafferenzprinzip
(Wechselwirkungen zwischen Zentralnervensystem und
Peripherie). Naturwissenschaften 37:464-476
von Stein St (1910) Schwindel (Autokinesis externa et interna).
Lessier, Leipzig
Voss H (1958) Zahl und Anordnung der Muskelspindeln in den
unteren Zungenbeinmuskeln, dem M. sternocleidomastoideus
und den Bauch- und tiefen Nackenmuskeln. Anat Anz
105:265-275
Wapner S, Werner H, Chandler KA (I 951) Experiments on sensorytonic field theory of perception. J Exp PsychoI42:341-345
Weeks V, Travelli J (1955) Postural vertigo due to trigger areas in
sterno-cleidomastoid muscle. J Pediatr 47:315-327
Weiland W (1912) Hals- und Labyrinthreflexe beim Kaninchen:
ihr Einflu auf den Muskeltonus und die Stellung der
Extremitten. Pflgers Arch Ges PhysioI147:1-27
Wilson VJ, Boyle R, Fukishima K, Rose PK, Shinoda Y, Sugiuchi Y,
Vchino Y (1995) The vestibulocollic reflex. J Vestib Res
5:147-170
Zangemeister WH, Stark L (1983) Pathological types of eye and
head gaze co ordination in neurological disorders. Neurol
OphthalmoI3:259-276
Zikmund V (1966) Oculomotor activity during visual imagery of
a moving stimulus pattern. Stud Psychol (Praha) 8:254-272

SECTION K
Psychogenic vertigo

Chapter

31

Psychiatrie disorders and vertigo

The sensation of vertigo, a subjective complaint, is


sometimes defined as a hallucination of movement
or an illusion. Since the evaluation of hallucinations
and illusions is an essential part of psychiatrie practice, such definitions place vertigo at the threshold of
clinical psychiatry (Trimble 1984). Psychogenic vertigo is relatively common among psychiatrie
patients, especially phobie postural vertigo (Chap.
32; p. 469), one of the three major p ychiatric conditions manifesting as psychogenic vertigo (Brandt
1996). Of 1370 consecutive neurologieal patients presenting at our dizziness unit in Munich between
1989 and 1995,15% had phobie postural vertigo.The
two other major psychiatrie conditions manifesting
as p ychogenic vertigo are panic disorder and agoraphobia, conditions that can exist separately or
together. According to the review of Furman and
Jacob (1997), the following guidelines are used in
current clinical practice for the diagnosis of psychiatrie dizziness: the dizziness is not characterised by
true vertigo, the dizziness can be replicated by
hyperventilation, psychiatrie symptoms precede the
onset of dizziness, and the dizziness occurs in
anxious or phobie individuals. The reliability, however, of these features is limited and thus Furman
and Jacob propose a more "narrow" definition of
psychiatrie dizziness as follows: the dizziness occurs
exclusively in combination with other symptoms as
part of a recognised psychiatrie symptom cluster
and this symptom cluster is not itself related to
vestibular dysfunction.
Dizziness is one of the 13 most frequent symptoms of pa nie attacks, but like vertigo it can also be a
somatic symptom in a number of psychiatrie conditions (Table 31.1), including schizophrenia, mood,
and somatoform and dissociative disorders. Except
in somatoform disorders where dizziness/vertigo
may be the most prominent manifestation, the other
conditions rarely occasion neurological evaluation
when there is an obvious primary psychiatrie
diagnosis.
To complicate matters, vestibular dysfunction frequently causes psychiatrie illness because of its

incapacitating nature. Schilder (1933) argued that


the importance of the vestibular apparatus and body
acceleration for the internal representation of Our
body image accounts for the seeondary psychiatrie
symptomatology in patients with primary vestibular
disorders. Although such a psychoanalytieal
eonceptualisation cannot be verified empirically,
the vestibular-psyehiatric interrelationships are
inescapable.
A considerable neurotic overlay may delay the
diagnosis of organic vertigo syndromes, e.g. in
patients with Meniere's disease. Conversely, a transient labyrinthine dysfunction, e.g. in vestibular
neuritis, may initiate development of neurosis, in
which anxiety, panie attaeks, and depression are
highly common. The neurosis then typieally preserves subjective vertigo and postural imbalanee,
Table 31.1. Forms of psychogenic vertigo and diuiness (aording to DSMIV
classifieationl
Verligo os an Ossocioled symptom:
(295.0)
Schizophreni,
(297.1)
Delusion,l (Paranoid) Disorder (Somatic type)
(296.2)
Major Depressive Disorder
(300A)
Dysthymic Disorder (01 Depr",sive Neurosis)
(300.02)
Generalised An'iety Disorder
(300.11)
Conversion Disorder (or Hysterical Neurosis Conversion Typel
(300.71
Hypochondriasis (ar Hypochondriacal Neurosis)
(300.81)
Somatoform Disorder
(300.6)
Depersonalisation Disorder (01 Depersonalisation Neurosis)
(300. I 9)
Factitlous Disorder with Physlcal Symptoms
(309.9)
Adjustment Disorders un!pecified (with Physical Complainl!)
Verrigo os 0 defined syndrome
(300.211
Pani< Disorder w,th Agoraphobia
(300.221
Agoraphobla wllhout History 01 Panic Disorder
(300.29)
Acrophobia (Speciflc Phobia)

Phobie Postural Verligo _____


______ nOI part of DSM IV
Spaee Phobia
Psychogenil overloy of orgonic verrigo syndromes

(V65.2)

From Brandt (1996).

455

Predlsposed Personalities
Mani!"'t Psychiatrie DIsorders
Deliberate Aggravation ofSymptoms
Malingering

Vertigo

456

despite the patient's recovery from the peripheral


lesion or its central compensation. In the latter case,
there is usually a history of an initially rotational
vertigo, lateropulsion, and nausea (along with the
objective dinieal signs of horizontal rotatory nystagmus and unilateral hyporesponsiveness to calorie
irrigation). A more indefinite chronic dizziness
ensues with subjective postural imbalance or spells
of vertigo during active body movements, which are
associated with anxiety and change in behaviour
(see Phobie postural vertigo, Chap. 32).
Obsessional subjects are more likely to proceed
from organie to psychiatrie disease, and the obvious
presence of an obsessional personality will support
the diagnosis of psychogenic vertigo. Patients with
somatisation symptoms consult a specialist about
those symptoms of primary concern; with vertigo, it
is the neurologist or otolaryngologist to whom they
turn. The frequency of psychogenic vertigo and the
associated diagnostie problems are well-known.

Organie versus psychiatrie morbidity


The evaluation of patients at medical and neurological
(not psychiatrie) dinics has indeed proven that
many patients with primary psychiatrie disease first
present with organic symptoms. This fact may not be
widely known in clinical practice. Psychiatrie morbidity has been estimated to be between 30% and
60% for inpatients and 50% and 80% for outpatients
attending medical (not psychiatrie) dinics (Lipowski
1967), and among these patients depression was the
most frequent diagnosis (Shevitz et al. 1976). Of 4470
consecutive neurological inpatients presenting with
"typical neurologie al symptoms", 405 (9%) had a
psychogenic cause rather than neurological dysfunction of the nervous system as the primary reason for
admission (Fig. 31.1) (Lempert et al. 1990). Pain
(including headache) was the most common psychogenic symptom, followed by motor symptoms (in
particular stance and gait disturbanees; p.461),
dizziness (phobie postural vertigo: 47; other psychogenic vertigo syndromes: 38), psychogenic
seizures, sensory symptoms, and visual dysfunction.
In a total of 2716 neurological outpatients in
England, primary psychiatrie disability was found to
exist in 13.2%; headache was the leading complaint
followed by vertigo as the second in 23% of the total
(Kirk and Saunders 1977). In these cases ofvertigo,
neuro ses and personality disorders predominated in
the psychiatrie diagnoses, whereas depression
occurred less frequently, with an incidence of only
17%. Rating scales to assess psychopathology
revealed psychiatrie morbidity in 50% of 126 neurological inpatients and cognitive dis orders in 30%
(De Paulo and Folstein 1978). Results were similar

20
15
10
5

10

20

30

40

50

60

70

80

age
Fig.31.1. Age distribution of 405 neurological inpatients with
psychogenic dysfunction as the chief complaint as compared to
2042 randomly selected neurological inpatients (shaded areal.
(From Lempert et al. 1990.)

(46% and 33%, respectively) in a sm aller number of


57 consecutive admissions to general medical wards
(Knights and Folstein 1977).

Vestibular dysfunction secondary to psychiatrie


disorders and psychiatrie disorders secondary
to vestibular dysfunction
Attempts to define a neurophysiological link
between psychiatrie dis orders and somatic symptoms have concentrated on vertigo for centuries (de
Sauvages 1770-1771; Benedikt 1870; Freud 1895)
even before the remarkable group of 19th century
scientists (Jan Evangelista Purkyne, Ernst Mach,
Tosef Breuer, Hermann Helmholtz, and Alexander
Crum-Brown) discovered the functions of the
labyrinth and laid the foundation of modern
vestibular and ocular motor research (Cohen 1984;
Crum-Brown 1874; Mach 1875; Grsser 1984; Brandt
1991 b). The vestibular-psychiatric interrelationship
is characterised by the following aspects (Brandt
1998):

Schizophrenie patients often have abnormal


responses on vestibular testing (Fitzgerald and
Stengel1945)
Patients with anxiety neuroses exhibit greater
sensitivity and directional preponderance in
vestibular testing (Hall pike et al. 1951) and have
abnormal posturographic test results (Yardley et
al. 1994)
Psychiatrie, in particular schizophrenie, patients
are more susceptible to motion sickness
(Mirabile and Glueck 1980)

457

Psychiatrie disorders and vertigo

Feeling dizzy or unsteady is among the 13 major


symptoms of panic attacks as defined in DSM-IV
(American Psychiatrie Association 1994), and the
dizziness may even be present between attacks
(Jacob et al. 1989)
Psychiatrie morbidity is high in unselected
patients with vestibular disorders (Sullivan et al.
1993) and remains high in strictly selected
patients with documented organic vestibular
dysfunction (Eagger et al. 1992)
Patients with vestibular symptoms, particularly
with Meniere's disease, have a high frequency of
abnormalities on psychometrie tests (Hinchcliffe
1967; Singerman et al. 1980; Rigatelli et al. 1984)
Organic vestibular disease may precipitate as
sequelae panic attacks with or without agoraphobia (Pratt and McKenzie 1958; Lilienfeld et al.
1988) and forms of psychogenic vertigo such as
"motorist disorientation syndrome" (Page and
Gresty 1985), space phobia (Marks 1981), and
phobie postural vertigo (Huppert et al. 1995).

subjects. Therefore, one should expect the abnormal


mental states in primary psychiatrie patients under
influence also of medication to bias the standardised
experimental conditions under which vestibular
tests, such as vestibulo-ocular reflex (VOR), fixation
suppression of the VOR or caloric irrigation, are performed. It has been shown that subjective anxiety is
correlated with an enhanced gain of the VOR, and
that stressful mental activity results in an increase in
maximum slow phase velo city of VOR response to
rotational testing (Yardley et al. 1995). Rotational
testing and caloric responses, however, show such
large interindividual variability that acceptable normal data cannot be established for practical clinical
use. A positive finding can be confirmed in an apparently normal individual only when there is a difference of at least 20% between the left- and right-side
responses.

The shortcomings of all these reports in the copious


literature have a considerable bearing on the management of vertigo patients. From the design of most
studies, however, we cannot expect the statistical
data to elucidate our current knowledge of the
neurophysiological mechanisms involved in the
interrelationship between lesional dysfunction and
psychiatrie manifestation. Some reports are purely
speculative or present psychoanalytic explanations,
ignoring sensory physiology; others are purely
phenomenological, comparing gross differences
between normal subjects and a heterogeneous group
of vertigo patients. This approach remains unsatisfactory for the clinician, because it does not help to
establish a particular diagnosis of peripheral
labyrinthine or central vestibular dysfunction, which
is clearly defined by a pattern of perceptual, oculomotor, and postural deficits. Moreover, most studies
lack a convincing neurophysiological hypothesis (for
example, the extensive literature on abnormal eye
movements in schizophrenics). If one examines a
number of vertigo patients consecutively admitted
to a clinic of either specialisation, one will find that a
considerable proportion of conditions are purely
psychosomatic. This is already evident from the frequency distribution of the diagnostic categories
(Table 3l.l) and thus significantly increases the
abnormal scores on personality tests. Furthermore,
if vestibular dysfunction, like pain, is accepted to
involve psychic reactions (in particular anxiety and
depression) then it is not surprising that vestibular
testing - dependent as it is on relaxation, attention,
and expectation - shows greater variability and
abnormal results than vestibular testing of healthy

The diagnostic approach to a classification of vertigo


syndromes is based mainly upon the patient's own
description of his symptoms. "Assessment of its aetiology, particularly where objectively determined
neurological accompanying signs are minimal or
absent, will depend therefore on an understanding of
the patient in whom the symptom is arising, his
background quality of interpersonal relationships,
and the life situation in which the symptoms have
arisen" (Trimble 1984). It is very important to
explore anxiety and depression, because patients
often tend to perceive them as areaction to rather
than the cause of the disease. In order to differentiate an organic vertigo syndrome in a psychiatrie
patient, an organic vertigo syndrome with secondary
neurotic overlay, and a purely psychosomatic vertigo
syndrome in an otherwise healthy patient, clinical
experience in both otoneurology and psychiatry is
required. The absence of pathology on otoneurological examination (as is sometimes the case, for exampie, in the fatigue phase of benign paroxysmal
positioning vertigo) of course does not prove that
the vertigo complaint is psychogenic, since obvious
psychiatrie disability does not exclude lesional
vestibular dysfunction.
Clinical experience allows us to formulate some
characteristics of psychogenic vertigo (Table 31.1) as
a guide to differential diagnosis.
Psychogenic vertigo should be suspected if

How can psychogenic vertigo be diagnosed?

certain stimuli or social events are the main


causes,
there is a clear dissociation between objective
and subjective disequilibrium,

Vertigo

458

the patient complains of a distinct rotational vertigo without concurrent spontaneous nystagmus
while wearing Frenzel's glasses,
the patient experiences inappropriately excessive
anxiety or fear of impending death.

Unusual features of psychogenic vertigo are

nausea and emesis, which are typically associated


with acute peripherallabyrinthine and vestibular
nuclei lesions,
rotational vertigo with direction-specific falls.

Some forms of otolith vertigo (p. 29) may present


without easily recognisable neurological signs.
Vertigo attacks lasting for a fr action of a second may
indicate psychogenic but also organic vertigo, such as
vestibular paroxysmia (see Chap. 7) or the "blip" syndrome (Lance 1996). The latter describes momentary
sensations of impending loss of consciousness, particularly when a person is relaxed without any obvious
cardiac, cerebral vascular, or epileptic basis. These
episodes may be a quasiepileptic phenomenon such
as deja vu and night starts, and may seem to have a
benign prognosis (Lance 1996).
Dizziness is the main clinical presentation of the
hyperventilation syndrome (p.23), regardless of
whether it is causative or secondary.
In anxiety neurosis with vertigo the patient typically complains of either a purely defined dizziness
and postural imbalance or brief attacks of vertigo.
Vertigo associated with depression presents as an
ill-defined lightheadedness or dizziness in conjunction with concentration difficulties and slowness of
thought. A masked depressive illness may manifest
purely somatically, but it can usually be recognised
by other depressive symptoms such as poor sleep,
loss of appetite and libido, retardation of activity, or
agitation and increased irritability as well as depressive anxiety. It should be noted here, after Trimble
(1984), "that hypochondriacal preoccupation with
vertigo often disguises an underlying depressive illness, in such patients the persistence and tenacity of
their complaints always seeming at odds with the
apparent negative findings of neurological investigation".
Hysterical vertigo, sometimes bizarre and combined with psychogenic astasia/abasia, is easily
recognised as a hysterical symptom if the patient's
personality assurnes a florid form with exhibitionism and histrionies. However, only approximately
one-fifth of patients who develop conversion
phenomena (vertigo may be one such presentation)
have such personality styles. This indicates that
symptoms of hysteria can occur in a variety of different personalities (Trimble 1984). Williams (1975)

particularly emphasises the obsessional personality


styles in such patients: "what they have in common
is innate stoicism, which not only enables them, but
compels them, to maintain control of events according to their own high standards in circumstances
where by ordinary criteria an overt emotional
response might be expected". It should be mentioned here that "hysterical appearance" may reflect
some previous or currently underlying vestibular
dysfunction. It has been reported that the majority
of patients presenting with hysterical symptoms had
a relevant co-existent or previous organic disease
(Whitlock 1967; Merskey and Buhrich 1975). The
best example of this is epilepsy.
The so-called post-concussion syndrome (see Posttraumatic vertigo, Chap. 22) includes headache, psychiatrie symptoms, dizziness, visual dis turban ce, and
fatigue, and is often attributed to neurosis. Since
post-traumatic vertigo symptoms in many patients
resemble those seen in otolith dysfunction, one can
speculate that this vulnerable accelerometer is
affected by traumatic loosening of otoconia, which
results in unequally heavy loads on the macula beds
and a vestibular tonus imbalance (Brandt and Daroff
1980). Thus, post-traumatic vertigo may be organic
vertigo of the otolith type (Brandt 1991 a; Gresty et
al. 1992), which goes undetected in routine clinical
testing. To further complicate matters, it may be
combined with, or initiate, neurotic development.
Careful assessment of patient history is required to
distinguish immediate symptomatology from the
neurotic complications that develop later.
Intoxication must be excluded along with iatrogenie vertigo. The latter may include posture and
gait instability in patients under medical treatment,
e.g. antiepileptic drugs (see Vertigo as adverse pharmacological effect, Chap. 28).
The management of purely psychogenic vertigo
consists of sympathetic counselling, designed to
relieve the patient of fear of organic disease, and
behavioural therapy, supported by short-term prescription of tricyclic antidepressants or tranquillisers
as the need arises.

Panic disorder
According to DSM-IV (American Psychiatrie
Association 1994), "the essential feature of Panic
Disorder is the presence of recurrent, unexpected
Panic Attacks followed by at least one month of persistent concern about having another Panic Attack,
worry about the possible implications or consequences of the Panic Attacks, or a significant

459

Psychiatrie disorders and vertigo

behavioral change related to the attacks (Criterion


A). The Panic Attacks are not due to the direct physiological effects of a substance (e.g. Caffeine
Intoxication) or a general medical condition (e.g.
hyperthyroidism) (Criterion C). Finally, the panic
attacks are not better accounted for by another
mental dis order (e.g. Specific or Social Phobia,
Obsessive-Compulsive Disorder, Post-traumatic
Stress Disorder, or Separation Anxiety Disorder)
(Criterion D). Depending on whether Criterion Ais
also met for Agoraphobia, Panic Disorder With
Agoraphobia or Panic Disorder Without
Agoraphobia is diagnosed (Criterion B)".
A panic attack is a discrete period in which there
is the sudden onset of intense apprehension, fearfulness, or terror, often associated with feelings of
impending doom. During these attacks, symptoms
such as shortness of breath, palpitations, ehest pain,
or discomfort, choking or smothering sensations,
and fear of "going crazy" or losing control are present. Panic disorder is not a distinct diagnostic entity,
and the overlap of panic dis order with other psychiatrie disorders raises the broader quest ion of
whether it is even a distinct psychiatrie dis order
(Aronson 1987).
Phobias present many facets and borderline variants as they progress. The point made by Sheehan et
al. (1980) is certainly correct, i.e. that the various
diagnostic labels usually refiect the diagnostic interests of the doctors who treat the patients.

Criteria for panic attack (DSM-IV 1994)


A discrete period of intense fear or discomfort in
which four (or more) of the following symptoms
develop abruptly and re ach a peak within 10 min:
1.
2.
3.
4.

5.
6.
7.
8.
9.
10.
11.
12.
l3.

palpitations, pounding heart, or accelerated


heart rate,
sweating,
trembling or shaking,
sensations of shortness ofbreath or smothering,
feeling of choking,
ehest pain or discomfort,
nausea or abdominal distress,
feeling dizzy, unsteady, lightheaded, or faint,
derealisation (feelings of unreality) or depersonalisation (being detached from one's self),
fe ar of losing control or going crazy,
fear of dying,
paraesthesias (numbness or tingling sensations),
chills or hot fiushes.

Agoraphobia
Since agoraphobia occurs in the context of Panic
Disorder With Agoraphobia and Agoraphobia
Without History of Panic Disorder, the criteria for
agoraphobia are provided separately in DSM-IV
(American Psychiatrie Association 1994).

Criteria for agoraphobia


"A. Anxiety about being in places or situations from
which escape might be difficult (or embarrassing) or
in which help may not be available in the event of
having an unexpected or situationally predisposed
Panic Attack or panic-like symptoms. Agoraphobie
fears typically involve characteristic clusters of situations that include being outside the horne alone;
being in a crowd or standing in a line; being on a
bridge; and traveling in a bus, train, or automobile.
(Note: consider the diagnosis of Specific Phobia if
the avoidance is limited to one or only a few specific
situations, e.g. height) or Social Phobia if the avoidan ce is limited to social situations.
B. The situations are avoided (e.g. travel is restricted)
or else are endured with marked distress or with
anxiety about having a Panic Attack or panic-like
symptoms, or require the presence of a companion.
C. The anxiety of phobie avoidance is not better
accounted for by another mental disorder, such as
Social Phobia (e.g. avoidance limited to social situations because of fear of embarrassment), Specific
Phobia (e.g. avoidance limited to a single situation
like elevators), Obsessive-Compulsive Disorder (e.g.
avoidance of dirt in someone with an obsession
ab out contamination), Post-traumatic Stress
Disorder (e.g. avoidance of stimuli associated with a
severe stressor), or Separation Anxiety Disorder (e.g.
avoidance of leaving horne or relatives)" (Criteria for
Agoraphobia, DSM-IV 1994).

Impaired balance was found in patients with panic


disorder with agoraphobia when proprioceptive balance information was minimised by sway-referencing of the support surface (Jacob et al. 1997).
Agoraphobies seem to rely on proprioceptive cues
for maintaining balance.
Epidemiology

Epidemiological studies indicate the lifetime prevalence of Panic Disorder With/Without Agoraphobia
to be between 1.5 and 3.5%; the age of onset is most
typically between late adolescence and the mid-thirties

Vertigo

460

with abimodal distribution (a second smaller peak in


the mid-thirties, especially in males) (Hand and
Wittchen 1986). Phobias are considered to be an
inherited illness (women/men: 2/1) with a lifelong
probability that first-degree relatives of patients with
manifest phobia will develop phobias themselves
(Crowe et al.1983). The risk in the case of agoraphobia
is estimated to be higher than 30% (Harris et al.
1983). Additional significant risk factors are being
female, young and poorly educated. Reports on
racial differences in the prevalence of phobie disorders indieate that prevalence is higher for blacks
than whites. These racial differences remain even
when demographic and socioeconomie factors are
held constant (Brown et al. 1990).
Panie disorder patients are also more likely than
the general population to have peptic ulcer disease,
hypertension, mitral valve prolapse, irritable bowel
syndrome, asthma, migraine, headaches, and a typical chest pain diagnosis (Stahl and Soefje 1995).
Twenty-seven percent of panic dis order patients will
abuse alcohol or drugs; more than 80% will experience depression (Wittchen and Essau 1993).
Management

Psychiatric treatment of panie dis order and agoraphobia is based on psychopharmacotherapy,


psychotherapy and combinations thereof (Stahl
and Soefje 1995; Frommberger et al. 1995).
Pharmacotherapy includes benzodiazepines, trieyclic
antidepressants, mono amine-oxidase inhibitors, and
selective serotonin-reuptake inhibitors. Psychotherapy
includes cognitive-behavioural therapy and psychodynamic psychotherapy. Both approaches have
proved their efficacy in reducing panic attacks and
agoraphobia. The better results achieved with combined treatment in the short run have not proven
superior to cognitive-behavioural therapy without
pharmacotherapy in the long run (Frommberger et
al. 1995). Cognitive-behavioural therapy is generally
considered the most effective therapy available for
panie dis order (Michelson and Marchione 1991). A
weak but statistically significant placebo effect has
been detected in agoraphobic patients (Mavissakalian
1988).
All these treatments have to be compared with the
natural course of panie dis order and phobias. The
long-term course of untreated phobias indicates that
during a 5-6 year interval, most children's phobias
and 40-60% of adult's phobias either recover or
improve substantially (Ag ras et al. 1972; Noyes et al.
1980). A follow-up study of panic dis order found
remissions in 43% of the patients; avoidance behaviour and chronie anxiety were more persistent than
panic attacks over a l-year period (Maier and Buller

1988). The latter clinieal studies, however, do not


exclude the possibility of later relapses.

Acrophobia
Acrophobia is a specific phobia characterised by significant anxiety provoked by exposure to height,
which often leads to an extreme avoidance behaviour. Neurotic acrophobia results when physiologieal
height vertigo (see Height vertigo, p. 418) induces a
conditioned phobic reaction which is characterised
bya dissociation between the objective and the subjective risk of falling (Brandt 1984). Although the
acrophobie is normallyaware of this dissociation, he
cannot overcome his avoidance behaviour. Also,
agoraphobia, as long as it mainly involves subjective
postural imbalance in wide open spaces, may be
initiated by the physiologieal impairment of visual
control of body sway consequent upon the increasing distance to stationary objects in the seen
environment.
Panic attacks in phobic vertigo syndromes
require the presence of both neurotic structure of
personality and the eliciting stimulus situation,
which is often unpleasant even for healthy subjects.
Consequently, impairment of postural balance, due
to ataxia or a deficiency in any of the stabilising sensory systems, may facilitate the induction of acrophobia or agoraphobia in predisposed subjects. "I
am afraid when 1 look down from a high place", the
MMPI (Minnesota Multiphasie Personality
Inventory) item 166, elicits a positive response from
the majority of alcoholics. Compared with the profiles of normal subjects and psychiatrie patients,
they have a significantly greater fear of heights
(Curlee and Stern 1973). Also, Pogany's (1958)
empirical report that patients with vestibular dysfunction exhibit a greater susceptibility to fear of
heights is now understandable, since the erroneous
visual signal should have a disproportionately
greater sensorial weight when associated with a
vestibular lesion. His second finding, however, of a
hearing defect in more than 90% of acrophobic
patients, was not confirmed by others, and its physiological basis remains unexplained. The occasional
development of acrophobia following head or
whiplash injury has been attributed to ophthalmologieal abnormalities, mostly eye muscle paresis
(Takeya et al. 1978), but it is more likely to occur as a
post-traumatie phobie postural vertigo (p. 469),
sometimes initiated by a traumatic lesion of the
otoliths (p. 349).
Neurotic fear of heights can be readily predieted

Psychiatrie disorders and vertigo

by use of general or specific psychological trait anxiety tests (e.g. State-Trait Anxiety Inventory; A-Trait
Scale; Taylor Manifest Anxiety Scale; Neuroticism
Scale of the Eysenck Personality Inventory; Geer
Fear Survey Schedule). The superiority of personality
questionnaires specifically designed for this purpose
was demonstrated by Mellstrom et al. (1976), who
compared test results and fear measures in actual
stimulus situations. Acrophobic questionnaires
based on self-report measures are more reliable than
laboratory-type behavioural tests (Cohen 1977).

Psychotherapy tor acrophobia and


agoraphobia
Psychotherapy of acrophobia (Table 31.2) is dominated by behavioural approaches which can be classified as either systematic or in vivo desensitisation
procedures. Goethe (1771) in the ninth book of the
Strassburger Tischgesellschaft, Selbsterziehung, gives
a description of how he treated his acrophobia by
climbing up daily to the top of the Strassburger
Mnster. In modern terms his technique can be
defined as a self-controlled in vivo desensitisation
somewhere between "successive approximation" and
"flooding".
The technique of systematic desensitisation after the earliest studies by Wolpe (1958) - is based
on the construction of a graduated hierarchy of
anxiety-provoking visual scenes, which after a training phase in muscular relaxation are subsequently
visualised by the patient while in the relaxed state.
Although therapist- or self-directed (tape-recorded)
desensitisation has been widely used (Baker et al.
1973; Takeya et al. 1979), it has been demonstrated
that in vivo desensitisation is more effective in the
treatment of phobia. Behavioural therapy for acrophobia can be combined with psychopharmacotherapy (p. 475; Zitrin et al. 1980).
In vivo desensitisation techniques aim to reduce
avoidance behaviour or anxiety by increasing the
patient's contact with reallife rather than with imagined provoking stimuli. Successive approximation to
the feared situation is supported by instructions and
reinforcement. Contact desensitisation after Ritter
Table 31.2. Behavioural therapy of acrophobia

Systematic desensitisation
- Therapist or self-direeted
-Implosion therapy

In vivo desensitisation
- Sueeessive approximation
- Contaet desensitisation
- Flooding

461

(1969a,b) stresses the advantage of joint participation and physical contact with the therapist-model
(participant modelling), during a graded approach
to heights. "Flooding", an alternative technique, also
uses exposure to real-phobie stimuli, but without a
graduated approach. It is based on getting the
patient into a feared situation and maintaining the
exposure for prolonged periods of time. Flooding in
imagination (implosion therapy) relies on the
description of high anxiety scenes and is less effective. The successful treatment of mentally retarded
patients suffering from acrophobia (Guralnick 1973;
Peck 1977) has proven that the seemingly implicit
demand for verbal skill in the techniques of desensitisation is not a necessary precondition.

Psychogenic disorders of stance and


gait
The most common psychogenic movement disorders are tremor, dystonia (Marsden 1995), and gait
disturbanees. They account for more than 75% of
psychogenic movements (Williams et al. 1995). For
the physician patients with psychogenic disorders of
stance and gait frequently appear to have balance
problems because they exhibit an increased or
bizarre postural sway with falls, near tumbles, or
gross stumbling, but these patients seldom complain
themselves about vertigo. As a rule, they complain in
a less differentiated way about aglobai stance or gait
disorder and leave it up to the physician to classify
their disorder clinically. Patients with psychogenic
disorders of stance and gait exhibit their obvious
dis ability, usually do not work, and sometimes seem
to suffer little. Among 405 patients presenting with
psychogenic disorders, 47 were diagnosed to have
phobie postural vertigo and 52, psychogenic disorders of stance and gait (Lempert et al. 1990).
Phobie postural vertigo and psychogenic dis orders
of stance and gait accounted for about 1% each of
the diagnoses in our inpatients.
The importance and frequency of psychogenic
dis orders of stance and gait are not reflected in the
reeent literat ure and textbooks. For general deseriptions, neurologists have to search through old textbooks or a few reeent surveys (Strmpell 1899;
Dejerine 1914; Janet 1920; Bing 1924; Keane 1989;
Lempert et al. 1991). We have tried to establish diagnostie eriteria for psyehogenie disorders of stanee
and gait (Lempert et al. 1991) as previous authors
have done for psychogenie seizures (Desai et al.
1982; Gulliek et al. 1982; Gates et al. 1985; Luther et
al. 1985). The point in question was whether diagnosis

462

Vertigo

is possible on phenomenological grounds alone and


whether an analysis of single features or a classification into characteristic sub types is more practical
for diagnosis. Thirty-seven patients with psychogenic disorders of stance and gait were clinically
evaluated, recorded on video, and analysed with
regard to clinieal phenomenology (Figs. 31.2-31.9).
Normal findings on neurologieal evaluation are a
precondition for the diagnosis of purely psychogenic
disorders of stance and gait. Psychiatrie findings are
heterogeneous in patients presenting with this condition and include conversion disorder and depression.
Six clinical features proved more valuable for the
diagnosis since they occurred alone or in combination in more than 90% of our tested patients
(Lempert et al. 1991; Table 31.3).

Criteria for psychogenic disorders of stance and


gait
1. momentary fluctuations of stance and gait, often

in response to suggestion
2. excessive slowness or hesitation of locomotion
incompatible with neurologieal disease
3. psychogenic Romberg test with a build-up of
sway amplitudes after latency or with improvement upon distraction
4. uneconomic postures with wastage of muscle
energy
5. the "walking on ice" gait pattern, which is characterised by small cautious steps with fixed ankle
joints,
6. sudden buckling of knees, usually without falls
Fluctuations of gait and stance performance incompatible with neurologieal diseases (Fig. 31.2)
occurred in more than half of our patients.
Encouragement and distraction by certain tasks,
Table 31.3. Charaeteristie features of psyehogenie dysfunetion of stanee
and gait in 37 patients

Fluetuation of impairment
Exeessive slowness of movements
Hesitation
"Psyehogenie Romberg" test
"Walking on iee" gait pattern
Uneconomie postures with waste of muscle energy
Sudden buekling
Without falls
With falls
Astasia
Vertieal shaking tremor
From Lempert et al. (1991).

19
13
6
12
11
11
10
8
2

4
3

such as finger-no se testing, considerably improved


the disturbance (Figs 31.3 and 31.6). Sometimes gait
returns to normal when walking is rendered more
difficult, for example, when walking on tip-toe.
Fluctuations mayaiso occur in neurologieal disease
such as myasthenia gravis, dyskalaemie paralysis,
and extra pyramidal disorders. Momentary to-andfro fluctuations, however, are indicative of a psychogenic origin.
Slow motion of gait (Fig. 31.4) is also very typieal
and is usually brought about by simultaneous innervation of antagonist muscles whieh may reflect the
confliet between voluntary intention and subconscious restraint. A related phenomenon likewise
caused by co-innervation of antagonist muscles is
hesitation. In these patients, initiation of intended
movement is either delayed or impossible. Small forward and backward movements of the leg may be
observed while the feet seem to stick to the ground.
In contrast to Parkinson's disease, the inhibition is
not overcome once the first step is taken, but recurs
with each step. An organic syndrome of gait-ignition
failure has been described with start and hesitation,
shuffling and freezing which, unlike parkinsonism,
may be due to frontal lobe disorders (Atchinson et
al. 1993). In the elderly, higher level gait disorders,
such as "frontal or senile" disorders of gait (Nutt et
al. 1993) and gait "apraxia" are the most relevant for
differential diagnosis (Chap. 27, p. 386).
A psychogenic "Romberg" test reveals constant
falls toward or away from the observer, irrespective
of position (Figs 31.3 and 31.6). Falls are usually
avoided by clutching the physician. Furthermore, the
large-amplitude body sway builds-up after a silent
latency over a few seconds and improves when the
patient is distracted, for example by asking hirn to
recognise numbers written with a finger on his skin
or by other manouevres (Fig. 31.6).
Uneconomic postures are adopted which waste
muscle energy in order to maintain balance (Fig.
31.2). Examples are assumption and maintenance of
postures with an excentric displacement of the centre of gravity, standing and walking with flexion of
the hip and knees, and slowing or arresting a knee
bent half-way down. Camptocormia, or functional
bent-back syndrome, is characterised by a 30-90
anterior flexion of the trunk associated with back
pain (Rockwood and Eiiert 1969; Soreff 1983; Sinel
and Eisenberg 1990; Fig. 31.9). Sudden buckling with
the knees may occur with and without falls (Fig.
31.5). The patients usually prevent themselves from
falling by activating the entire antigravity muscles
before they touch the ground, which requires a considerable expenditure of strength. Even if falls occur,
they are more or Iess directed to avoid seIf-injury,
which makes them easily distinguishable from drop
attacks (p. 389).

Psychiatrie disorders and vertigo

"Walking on ice," a characteristic gait pattern


used by normals on slippery ground, consists of cautious, usually broad-based steps with decreased
stride lengths and height, stiffening of knees and
ankles, some degree of antagonist innervation and
shuffling of feet (Fig. 31.4). A gait abnormality in
patients with normal-pressure hydrocephalus shares
some of these features, lacks the anxious balancing,
but sometimes shows abduction of the arms, which
is characteristic of these patients.
Astasia is a psychogenic inability to stand without
support but with preserved function of the lower
limb in the recumbent position.
Additional signs suggesting a psychogenic disorder occurred in 73% of our patients (Lempert et al.
1991) (Table 31.4; Fig. 31.8). We called them "suggestive" since differentiation from neurological dysfunction was sometimes not as obvious as in the
characteristic features described above.
Pseudoataxia refers to an instability of posture
and gait characterised by demonstrative balancings,
sometimes with flailing of the arms and bizarre
excursions of the trunk, while the legs are often
unaffected. Patients may perform sudden sidesteps or tumble about, but the broad-based gait of
neurological ataxia is rare. In psychogenic monoparesis the affected leg is often dragged behind or
may be bizarrely twisted, usually in an equinovarus
position (Fig. 31.7). Temporary innervation of the
paralysed leg may be observed. Continuous flexion
or extension of the toes may be seen in patients with
otherwise normal muscle tone.
Psychogenic tremors - often resting, postural, and
kinetic - can be identified by their unusual distribution and their alleviation by distraction. In contrast
to most organic tremors, the distal portions of the
limbs may be spared and move limply in response to
proximal muscle activity.
During the study we had the clinical impression
that psychogenic disorders of stance and gait may be
classified into sub types according to predominant
features. The similarity of manifestations is striking
in some patients. The overall variability, however, is
large, and classification into characteristic subtypes
was not found useful (Table 31.5), because
predominant features varied from patient to patient
and occurred in various combinations. The use of
posturographic measurements has been proposed as
a means to differentiate malingerers from vertiginous patients, but the sensitivity and specificity of this
measure are limited and clearly inferior to the clinically experienced eye (Uimonen et al. 1995; Allum et
al. 1994). Allum et al. (l996a) investigated the alterations caused by active, potentially destabilising postural movements to normal balance correcting
responses when healthy subjects pretended to be
unstable. On the basis of their observations, a classification procedure was developed to objectively

463
Table 31.4.

Suggestive features of psychogenic dysfunction of stance


and gait in 37 patients
n

Motor symptoms (n = 22)

Pseudoataxia
Of the legs
Ofthe trunk
Sudden sidesteps
Flailing of the arms
Dragging of the legs
Continuous flexion of the toes
Continuous extension ofthe toes
Bizarre tremors (hands 7, legs 3, trunk 2, head 1)

3
6
6
3
4

2
1
9

Expressive behaviour (n = 19)

Grasping the leg


Mannered posture of hands
Suffering or strained facial expression
Moaning
Hyperventilation

6
6

15
8

From Lempert et al. (1991).

Classification of psychogenic dysfunction of stance and gait


into characteristic subtypes (37 patients)

Table 31.S.

Siow gait
Astasia
Pseudoataxia
Buckling

6 (16%)

Othertypes

12a (31 %)

3 (8%)
3 (8%)

Combinations

11

(30%)

2 (6%)

'Includes pseudomyoclonus, rhythmical side-to-side rocking, fixed flexion


of hips and knees, generalized tremor, "walking on ice'; tripping propulsion
with falls, hesitation, monoparesis with dragging of the leg, unilateral
limp, bilateral equinovarus posture of the feet, waddling in a squatting
position, dragging along (like an exhausted wanderer in the desert)
From Lempert et al. (1991).

identify the responses of patients with non-organic


"vertigo and dysequilibrium" to dynamic posturography (Allum et al. 1996b).
In conclusion, we advocate the phenomenological
approach to psychogenic dysfunction of stance and
gait, since it is direct, effective, and avoids unnecessary,
costly, and potentially harmful diagnostic procedures.
Management

Our therapeutic regimen is symptom-oriented and


aims at improving stance and gait by using positive
reinforcement supported by placebo infusions and
physical therapy. Most patients do not request further
psychotherapy on ce they have regained postural stability. There is a negative correlation of short-term
improvement with symptom duration: in our followup study on 17 patients with psychogenic dis orders of
stance and gait the condition proved intractable in
five patients in whom it had lasted longer than 2 years
(Brandt et al. 1994; see also p. 476).

464

Vertigo

31.2.

31.3.
Fig. 31.2. Fluctuating gait disturbance with
waste of muscular energy: patient starts walking
upright, then goes slowly into a full knee bend
and becomes erect again on her way back. (From
Lempert et al. 1991.)

Fig. 31.3. Psychogenic instability af stance


with "nan-neuralogical" stiff leaning ta the side;
this ceases once the patient is distracted by finger-nase testing. (From Lempert et al. 1991.)

Fig.31.4.

"Walking on ice": hesitant, slow gait


with anxious balancing, stiff knees and ankles,
and reduced step height and stride length. (From
Lempert et al. 1991.)

31.4.

465

Psychiatrie disorders and vertigo

31.5.

31.6.
Fig.31.5. Psyehogenie buckling. Patient falls forward with
sudden flexion of hips and knees and remains in an uneeonomie half-flexed posture. (From Lempert et al. 1991.)

Fig.31.6. Psyehogenie pseudoataxic instability of stanee and


gait with wobbling of the trunk, bizarre balancing, and flailing
of the arms, which eeases on distraetion. (From Lempert et al.
1991.)

Fig. 31.7. Psyehogenic monoparesis of the right leg.


a Dragging of the twisted leg, grasping the leg. b "Paralysed"
right leg supports the body during swing phase of the left leg.
e Innervation of the "paralysed"leg after encouragement. (From
Lempert et al. 1991.)

a
31.7.

466

Vertigo

References

Fig.31.8. Psychomotor symptoms of psychogenic gait disorders: suffering facial expression and grasping of the leg. (From
Lempert et al. 1991.)

Fig.31.9. Functional bent-back syndrome (camptocormia) in a


62-year-old patient presenting with about 30 anterior flexion of
the trunk during stance and gait.

Agras WS, Chapin HN, Oliveau DC (1972) The natural history of


phobia. Arch Gen Psychiatry 26:315-317
Allum jHj, Huwilen M, Honegger F (1994) Objective measures of
non-organic vertigo using dynamic posturography. In: Taguchi
K, Igarashi M, Moris S (eds) Vestibular and neural front.
Elsevier, Amsterdam, pp 51-55
Allum jHj, Huwiler M, Honegger F (l996a) Prior intention to
mimic a balance dis order: does central set influence normal
balance corresponding responses? Gait Posture 4:39-51
Allum jHj, Huwiler M, Honegger F (l996b) Identifying cases of
non -organic vertigo using dynamic posturography. Gait
Posture 4:52-61
American Psychiatrie Association (1994) Diagnostic and statistical manual of mental disorder, 4th edn (DSM-IV). American
Psychiatrie Association, Washington DC
Aronson TA (1987) Is panic disorder a distinct diagnostic entity?
A critical review of the borders of a syndrome. j Nerv Ment
Disease 175: 584-594
Atchinson PR, Thompson PD, Frackowiak RSj, Marsden CD
(1993) The syndrome of gait ignition failure: Areport of six
cases. Movement Dis 8:285-292
Baker BL, Cohen DC, Saunders jT (1973) Self-directed desensitisation for acrophobia. Behav Res Ther 11 :79-89
Benedikt N (1870) "Platzschwindel". Allg Wiener Med Ztg
15:488-489
Bing R (1924) Lehrbuch der Nervenkrankheiten, 3rd edn. Urban &
Schwarzenberg, Berlin, pp 663-667
Brandt Th (1984) Visual vertigo and acrophobia. In: Dix MR,
Hood ID (eds) Vertigo. Wiley, Chichester, pp 439-466
Brandt Th (l991a) Vertigo: Its multisensory syndromes. Springer,
London, pp 291-304
Brandt Th (l991b) Man in motion. Historical and clinical aspects
ofvestibular function. Brain 114:2159-2174
Brandt Th (1996) Phobie postural vertigo. Neurology
46:1515-1519
Brandt Th (1998) Neuro-otological and psychiatrie abnormalities.
j Neurol Neurosurg Psychiatry 65:619
Brandt Th, Daroff RB (1980) The multisensory physiological and
pathological vertigo syndromes.Ann NeuroI7:195-203
Brandt Th, Huppert D, Dieterich M (1994) Phobie postural vertigo, a
first follow-up. j NeuroI241:191-195
Brown DR, Eaton WW, Sussman L (1990) Racial differences in
prevalence of phobie disorders. j Nerv Ment Dis 178:434-441
Cohen B (1984) The roots ofvestibular and oculomotor research.
Intro Hum Neurobiol3:121
Cohen DC (1977) Comparison of self-report and overt behavioural
procedures for assessing acrophobia. Behav Ther 8: 17 -23
Crowe RR, Noyes R, Pauls DL, Slymen D (1983) A family study of
panic disorder. Arch Gen Psychiatry 40: 1065-1 069
Crum-Brown A (1874) On the sense of rotation and the anatomy
and physiology of the semicircular canals of the internal ear. j
Anat Physiol 8:327-331
Curlee j, Stern H (1973) The fear of heights among alcoholics. Bull
Menniger CI in 37:615-623
Dejerine j (1914) Semiologie des affections du systeme nerveux.
Masson, Paris, pp 541-549
De Paulo jR, Folstein MF (1978) Psychiatrie disturbances in
neurological patients. Ann NeuroI4:225-228
Desai BT, Porter Rj, Penry jK (1982) Psyehogenic seizures. Arch
NeuroI39:202-209
de Sauvages F (1770-1771) Nosologie Methodique. Herrissant,
Paris
Eagger S, Luxon LM, Davies RA, Coelho A, Ron MA (1992)
Psychiatrie morbidity in patients with peripheral vestibular

Psychiatrie disorders and vertigo


disorder: Clinical and neuro-otological study. J Neurol
Neurosurg Psychiatry 55:383-387
Fitzgerald G, Stengel E (1945) Vestibular reactivity to caloric stimulation in schizophrenics. J Ment Sei 91:93-100
Freud S (1895) ber die Berechtigung, von der Neurasthenie
einen bestimmten Symptomencomplex als "Angstneurose"
abzutrennen. Neurol CentralbI12:50-66
Frommberger U, Angenendt J, Berger M (1995) Die Behandlung
von Panikstrungen und Agoraphobien. Nervenarzt 66: 173-186
Furman JM, Jacob RG (1997) Psychiatric dizziness. Neurology
48:1161-1166
Gates JR, Ramani V, Whalen S, Loewenson R (1985) letal characteristics of pseudoseizures. Arch NeuroI42:1183-1187
Goethe JW v (1771) In: Selbstbiographische Schriften, Dichtung
und Wahrheit. 9. Buch der Straburger Tischgesellschaft,
Selbsterziehung
Gresty MA, Bronstein AM, Brandt Th, Dieterich M (1992)
Neurology of the otolith function: Peripheral and central disorders. Brain 155:647-673
Grsser 0- J (1984) J.E. Purkyne's contributions to the physiology
of the visual, the vestibular and the oculomotor systems. Hum
NeurobioI3:129-144
Gullick TA, Spinks IP, King DW (1982) Pseudoseizures: ictal
phenomena. Neurology 32:24-30
Guralnick MJ (1973) Behaviour therapy with acrophobic mentally
retarded young adults. J Behav Ther Exp PsychoI4:263-265
Hallpike CS, Harrison MS, Slater E (1951) Abnormalities of the
caloric test results in certain varieties of mental dis order. Acta
OtolaryngoI39:151-159
Hand I, Wittchen HU (1986) Panic and phobias. Springer, Berlin,
Heidelberg, New York
Harris EL, Noyes R, Crowe RR, Chandry DR (1983) Family study
of agoraphobia. Arch Gen Psychiatry 40:1061-1064
Hinchcliffe ER (1967) Personality profile in Meniere's disease. J
LaryngoIOtolaryngoI81:476-481
Huppert D, Kunihiro T, Brandt Th (1995) Phobic postural vertigo
(154 patients): its assoeiation with vestibular disorders. J Audiol
Med 4:97-103
Jacob RG, Lilienfield SO, Furman JMR, Durrant JD, Turner SM
(1989) Panic dis order with vestibular dinical observation and
descriptions of space and motion phobic stimuli. J Anxiety Dis
3:117-130
Jacob RG, Furman JM, Durrant JD, Turner SM (1997) Surface
dependence: a balance control strategy in panic disorder with
agoraphobia. Psychosomatic Med 59:323-330
Janet P (1920) The major symptoms of hysteria, 2nd edn.
Macmillan, New York
Keane JR (1989) Hysterical gait disorders: 60 cases. Neurology
39:586-589
Kirk C, Saunders M (1977) Primary psychiatric illness in a neurological outpatient department in North-East England. Acta
Psychiatr Scand 56:294-302
Knights EB, Folstein MF (1977) Unsuspected emotional and cognitive disturbances in medical patients. Ann Intern Med
89:723-724
Lance JW (1996) Transient sensations of impending loss of consciousness: the "blip" syndrome. J Neurol Neurosurg Psychiatry
60:437-438
Lempert Th, Dieterich M, Huppert D, Brandt Th (1990)
Psychogenic disorders in neurology: frequency and dinical
spectrum.Acta Neurol Scand 82:335-340
Lempert T, Brandt T, Dieterich M, Huppert D (1991) How to identify psychogenic disorders of stance and gait. A video study in
37 patients. J NeuroI238:140-146
Lilienfeld SO, Jacob RG, Furman JMR (1988) Vestibular dysfunction followed by panic disorder with agoraphobia. J Nerv Ment
Dis 177:700-701
Lipowski ZJ (1967) Review of consultation. Psychiatry and psy-

467
chosomatic medicine. 11. Clinical aspects. Psychosom Med
29:20L-244
Luther JS, McNamara JO, Carwile S, Miller P, Hope V (1982)
Pseudoepileptic seizures: methods and video analysis to aid
diagnosis. Ann NeuroI12:458-462
Mach E (1875) Grundlinien der Lehre von den
Bewegungsempfindungen. Engelmann, Leipzig
Maier W, Buller R (1988) One-year follow-up of panic disorders.
Eur Arch Psychiatr Neurol Sei 238: 105-109
Marks IM (1981) Space "phobia": a pseudo-agoraphobic syndrome. J Neurol Neurosurg Psychiatry 44:387-391
Marsden CD (1995) Psychogenic problems associated with
dystonia. In: Wiener WJ, Lang AE (eds) Behavioral neurology
movement disorders. Raven Press, New York, pp 319-326
Mavissakalian M (1988) The placebo effect in agoraphobia -II. J
Nerv Ment Dis 176:446-448
Mellstrom M, Cicala GA, Zuckerman M (1976) General versus
specific trait anxiety measures in the prediction of fear of
snakes, heights and darkness. J Consult Clin PsychoI44:83-91
Merskey H, Buhrich NA (1975) Hysteria and organic brain
disease. Br J Med PsychoI48:359-376
Michelson LK, Marchione K (1991) Behavioral, cognitive, and
pharmacological treatments of panic disorder with agoraphobia: Critique and synthesis. J Consult Clin Psychol
59:100-114
Mirabile CS, Glueck BC (1980) Motion sickness susceptibility and
patterns of psychotic illness. Arch Gen Psychiatry 37:42-50
Noyes R, Clancy J, Hoenk PR, Hymen DJ (1980) The prognosis of
anxiety neurosis. Arch Gen Psychiatry 37:173-178
Nutt JG, Marsden CD, Thompson PD (1993) Human walking and
higher-level gait disorders, particularly in the elderly.
Neurology 43: 268-279
Page MGR, Gresty MA (1985) Motorist's vestibular disorientation
syndrome. J Neurol Neurosurg Psychiatry 84:729-735
Peck CL (1977) Desensitization for the treatment of fear in the
high level adult retardate. Behav Res Ther 15:137-148
Pogany E (1958) ber den Hhenschwindel. Mschr Ohrenheilk
Laryng-Rhinol92:209-213
Pratt RTC, McKenzie W (1958) Anxiety states following vestibular
disorders. Lancet II:347-349
Rigatelli M, Casolari L, Bergamini G, Guidetti G (1984)
Psychosomatic study of sixty patients with vertigo. Psychother
Psychosom 41:91-99
Ritter B (1969a) Treatment of acrophobia with contact desensitization. Behav Res Ther 7:41-45
Ritter B (1969b) The use of contact desensitization, demonstration-plus-participation and demonstration-alone in the treatment of acrophobia. Behav Res Ther 7:157-164
Rockwood CA, Eilert (1969) Camptocormia. J Bone Joint Surg
51:553-556
Schilder P (1933) The vestibular apparatus in neurosis and psychosis. J Nerv Ment Dis 78:1-23
Sheehan DV, Ballenger J, Jacobsen G (1980) Treatment of
endogenous anxiety with phobic hysterical and hypocondriacal
symptoms. Arch Gen Psychiatry 37:51-59
Shevitz SA, Silberfarb PM, Lipowski ZJ (1976) Psychiatrie consultations in a general hospital. Areport on 1000 referrals. Dis
Nerv Syst 37:259-300
Sine! M, Eisenberg MS (1990) Two unusual gait disturbances:
astasia, abasia, and camptocormia. Arch Phys Med Rehab
71:1078-1080
Singerman B, Rieder E, Folstein LM (1980) Emotional disturbance
in hearing dinic patients. Br J Psychiatry 137:58-62
Soreff H (1983) Camptocormia. Arch Orthop Trauma Surg
101:151-152
Stahl SM, Soefje S (1995) Panic attacks and panic disorder: The
great neurologie impostors. Semin NeuroI15:126-132
Strmpell A (1899) Lehrbuch der speziellen Pathologie und

468
Therapie der Inneren Krankheiten. vol. 3, 12th edn. Vogel,
Leipzig, pp 614-616
Sullivan M, Clark MR, Katon Wj, Fischi M, Russo D, Dobie RA,
Vorhees R (1993) Psychiatrie and otologie diagnoses in patients
complaining of dizziness. Arch Intern Med 153: 1479-1484
Takeya T, Baron JB, Ohno Y, Agathon M, Andersson JC, Ushio N,
Bessineton JC, Pacifici M, Lemaire V, Chavannes N, Noto R
(1978) Comparative study of post-traumatic and psychogenic
acrophobia (fear of height). Agressologie 19: 19-92
Takeya T, Ohno Y, Matsubara H, Yasuda K, Wantanabe S, Shinzato
R, Tanaka Y, Noda S (1979) Physiological changes in the treatment of acrophobia (fear of height). Clin OtolaryngoI4:197-205
Trimble MR (1984) Psychiatrie aspects of vertigo. In: Dix MR,
Hood DD (eds) Vertigo. Wiley, Chichester, pp 345-358
Uimonen S, Laitakari K, Kiukaaniemi K, Sorri M (1995) Does posturography differentiate malingerers from vertiginous patients?
JVestib Res 5:117-124
Whitlock FA (1967) The aetiology of schizophrenia. Acta
Psychiatr Scand 43:144-162

Vertigo
Williams D (1975) The borderland of epilepsy revisited. Brain
98:1-12
Williams DT, Ford B, Fahn S (1995) Phenomenology and psychopathology related to pyschogenic movement dis orders. In:
Wiener WJ, Lang AE (eds) Behavioral neurology movement disorders. Raven Press, NewYork, pp. 231-257
Wittchen H, Essau CA (1993) Epidemiology of panic disorder:
progress and unresolved issues. J Psychiatr Res 27:47-68
Wolpe J (1958) Psychotherapy by reciprocal inhibition. Stanford
University Press, Stanford
Yardley L, Luxon L, Bird J, Lear S, Britton J (1994) Vestibular and
posturographic test results in people with symptoms of panie
and agoraphobia. J Audiol Med 3:48-65
Yardley L, Watson S, Britton J, Lear S, Bird J (1995) Effeets of
anxiety arousal and mental stress on the vestibulo-ocular
reflex. Acta Otolaryngol (Stoekh) 115:597-602
Zitrin CM, Klein DF, Woerner MG (1980) Treatment of acrophobia
with group exposure in vivo and imipramine. Areh Gen
Psyehiatry 37:63-72

Phobie postural vertigo

Phobie postural vertigo (PPV) (Brandt and Dieterich


1986; Brandt 1996) has been described as a syndrome that is distinguishable from panic disorder,
agoraphobia, and the pseudo-agoraphobia syndrome "space phobia" (Marks 1981). Since we began
diagnosing it as a clinical entity, PPV has become the
second-most common cause of vertigo in our dizziness unit (p. 23). Closely connected with gait and
posture, PPV is characterised by a combination of
non-rotational vertigo with subjective postural
imbalance and unsteadiness, normal neuro-otological tests, and phobie avoidance behaviour mainly in
patients with an obsessive-compulsive personality.
The monosymptomatic subjective disturbance of
balance manifests with superimposed attacks that
occur with and without recognisable provoking factors in one and the same patient. The patient experiences them with and without accompanying excess
anxiety, misleading both patient and physician to
make a false diagnosis of organic disease. Although
an association of PPV with anxiety disorders is
evident, not a11 patients present with symptoms of
anxiety or panic during attacks of vertigo. However,
most patients develop a disabling "phobic/avoidance
pattern" (Kapfhammer et al. 1997). These patients
contact neurologists or otolaryngologists first and
foremost and not psychiatrists or psychotherapists,
since their prevailing complaint is distressing postural imbalance and vertigo.
The illusory perception of imbalanee and vertigo
can be explained by the hypothesis that the space
constancy meehanism in these patients is impaired,
resulting in a partial deeoupling of the efference
eopy for active head and body movements. The
severely incapacitating effeet of derangement of
their multimodal spatial orientation, movement pereeption, postural regulation, and control movement
in the gravitational field makes it alm ost inevitable
that vertigo should trigger panie attacks and avoidanee behaviour. The suseeptibility of these symptoms to disturbance makes vertigo an almost
invariant eoneomitant of panic attacks.

469

The desire of patients with PPV to learn about the


psyehogenie meehanisms and their willingness to
engage actively in treatment are encouraging experiences for physicians. Having studied and treated
PPV for more than 10 years, we consider this syndrome a distinct clinical entity with typieal features,
reliable diagnosis, and a generally favourable
response to short-term suggestive and behavioural
therapy (Brandt et a1. 1994).

The clinical syndrome


PPV occurs chiefly in patients with an obsessivecompulsive or narcissistic personality. The diagnosis
is primarily based on six charaeteristic features
(Brandt et a1. 1994):
1. Dizziness and subjeetive disturbance of balance

2.
3.

4.
5.

while standing or walking, despite normal


clinical balance tests such as Romberg, tandem
walking, balancing on one foot, and routine
posturogra phy.
Fluctuating, continuous unsteadiness or episodes
lasting seeonds to minutes or momentary pereeptions of i1lusory body perturbations.
Although the attaeks can oecur spontaneously,
there is usually a perceptual stimulus (bridge,
staircase, empty room, street) or social situation
(department store, crowd, restaurant, concert)
from which the patients have difficulty withdrawing and which they reeognise as provoking
factors. There is a tendency for rapid conditioning, generalisation, and avoidance behaviour to
develop.
Anxiety (57%) and distressing vegetative symptoms occur during or after vertigo. Most patients
have attaeks both with and without anxiety.
Obsessive-compulsive type personality, labile
affect, mild depression.

Vertigo

470

6. Onset of the condition frequently follows a


period of particular emotional stress (Table
32.1), a serious illness, or an organic vestibular
dis order.
PPV is not generally connected with a previous or
current vestibular pathology as others have assumed
(Furman and Jacob 1997; Bronstein et al. 1997). The
patients feel physically ill and complain of vertigo
and a disturbance of balance; their anxiety can only
be brought to light by asking appropriate questions.
Patients describe the vertigo as either a threatening
fluctuation of awareness combined with unsteadiness or a perception of illusory body motion, wh ich
is often present for mere fractions of seconds.
Sometimes unsteadiness and vertigo are prolonged
and seem menacing, the sufferer being terrified of
the possibility of an unexpected fall, a loss of consciousness, or severe impairment of body function.
Subjective postural imbalance is described as a feeling of weakness, an inhibition of movement or disturbed co ordination of the legs, head, and trunk.
Fluctuations of head position when standing, a
result of active body movements such as turning the
head or shifting weight from one foot to the other,
provoke unpleasant illusions of body acceleration
with the simuItaneous illusory movement of a
stationary environment. In mild cases the patients
try to find support or lean on something. Typical but
not obligatory features of the attacks include anxiety,
psychomotor restlessness with escape reactions, a
sudden desire to flee from the place where the attack
was provoked, aimless walking, or if seated, a rigid
grasping of the arms of the chair.
Panic and subsiding anxiety may sometimes lead
the patients to believe that they can avert a physical
catastrophe only by running away or adapting other
types of behaviour (taking tablets, for example).
Afterwards, however, normal day-to-day activities
are resumed with astonishing speed. Phobie vertigo
patients can, for instance, conceal an attack of vertigo
from their tennis partner on the way to the court,
despite considerable inner tension, by stopping, supporting themselves, and making small talk with the
partner - and then go on to playa normal tournament match. This discrepancy between the initial
panic attack and the patient's subsequent disregarding it typifies PPV, but is rarely observed in organic
forms of vertigo.
The attacks occur at irregular intervals, sometimes several times a day. Specific questioning
almost invariably evokes particular constellations of
perceptual stimuli or social situations, which the
patient recognises as provoking factors. But "spontaneous attacks" also occur, so that the patient usually
tends to ass urne that the cause is organic. After

repeated attacks a conditioning and generalisation


seems to take place, to the extent that everyday but
unexpected situations involving movement stimuli
may provoke abortive attacks without significant
accompanying anxiety. These include, for instance,
slow sweeping movements of a curtain in the
peripheral visual field, which induce an optokinetic
postural response (p. 424), or incorrect estimation of
the height of a step of stair, or stepping onto a soft
surface. Both of the latter cases require abrief readjustment of sensory motor programmes for postural
regulation. Patients with PPV are inclined to experience external stimuli of this type primarily as disorders of body function; they then often spend
considerable time in retrospective analysis of the
attacks to determine the provocation, and will check
their sense of balance and other brain functions
several times a day.
Anticipatory anxiety, which is rarely expressed
spontaneously by the patient, paves the way for further attacks despite the discrepancy between the
subjective fe ar of falling and the absence of objective
unsteadiness. On the one hand, patients may spontaneously volunteer the observation that there has
been a temporary improvement after drinking
alcohol or during periods of particular physical
activity. On the other, sometimes the attacks are
exacerbated after sleep withdrawal or the day following increased alcohol consumption. The longer the
symptoms persist, the easier it becomes to identify
the particular situations that act as stimuli. The
patients tend to avoid these situations increasingly,
and thus may become socially isolated and perform
less weIl at work. The designation "phobie" is
theoretically questionable, since PPV may occur
without concomitant anxiety (Furman and Jacob
1997; Eckhardt-Henn et al. 1997). Avoidance of
potentially precipitating stimulus situations is still a
major feature of the condition, even if it is not associated with major anxiety (Brandt et al. 1997). To our
surprise, most of our patients remained socially weIl
integrated and were able to continue their normal
work despite symptoms that they feit dominated
their lives. PPV patients tend to hide their problem.

Aetiology and hypothetical


mechanism
Huppert et al. (1995) performed a retrospective
studyon 154 patients diagnosed with PPV. The age
distribution ranged from 18-70 years with a clear
peak in the fourth and fifth decades (Fig. 32.1). In 32
patients (21 %), PPV foIlowed an acute vestibular dis-

471

Phobie postural vertigo

.,..ca
J!l
c

50

= 154

14 (44%)

40

c-

30

$J

E
:::l
C

20
10

10 - 20 - 30 - 40 - 50 - 60 - 70 - 80
age

Distribution of age and sex (F, open columns; M, fi/led


of 154 patients with phobie postural vertigo (PPV).
(From Huppert et al. 1995.)
Fig.32.1.
columns)

order, in particular benign paroxysmal positioning


vertigo (14 patients) and vestibular neuritis (10
patients). Most showed a transition from organic
vertigo to PPV without a symptom-free interval
(Fig. 32.2). Thus, in those at risk, a vestibular disorder may trigger PPV, which then takes a course
independently of vestibular recovery or compensation. The 21 % may appear to be even unexpectedly
low, but the percentage increases if patients are
included whose condition was initiated by head or
whiplash trauma. Symptoms of post-traumatic dizziness often resemble those of otolith dysfunction
(p. 29). It is tempting to speculate that the otoconia
are dislodged by the trauma; this causes unequal
loads on the macular bed which results in a tonus
imbalance between the two otoliths: otolithic vertigo
(p. 29; Brandt 1997; Brandt and Daroff 1980; Gresty
et al. 1992). The symptoms of otolithic vertigo
(which is cent rally compensated within weeks) are
similar to those of PPV, suggesting that the latter
evolves from an initial otolith dysfunction. A striking number of cases of PPV occur after aperiod of
psychosocial stressors (Table 32.1; Kapfhammer et
al. 1997) or a somatic illness.

Hypothetical mechanism: Adisturbance of


space constancy due to decoupling of the
efference-copy signal?
We know little about the subtle, self-generated body
ftuctuations or involuntary head movements that
occur during upright stance. The environment also
appears to be stationary during active movements,
although relative movements between head and surroundings cause shifts of the retinal image. "Space
constancy" seems to be maintained as follows: the
voluntary impulse for initiating movement is simul-

BPPV

Vestibular
Neuritis

Meniere's
Disease

Central Vestibular
Disorder

Basilar Artery
Migraine

Fig.32.2. Vestibular disorders assoeiated with phobie postural


vertigo (PPV) in 32 of 154 patients. Fi/led columns: dose temporal
relationship with the vestibular disorder direetly preeeding PPV;
open columns: no dose relationship. (From Huppert et al. 1995.)

taneously accompanied by an appropriate efferencecopy signal to make identification of self-motion


possible (Fig. 32.3). Von Holst and Mittelstaedt
(1950) suggested that the efference copy is used to
readjust the perceptual systems (based on previous
experience), in order to interpret incoming sensory
information that results from movement of the
observer relative to a stationary environment rather
than vice versa. A decoupling would cause the sensory effects of normal postural adjustments (of
which we are usually unaware) to be interpreted as
arising from either external body perturbations or
motion of the surroundings. If there is no efference
copy for body motion, e.g. if we move our eyeball by
placing the finger on the eyelid, we see illusory
movements of the environment, e.g. oscillopsia
(p.430). The sensation of vertigo described by
phobie patients (involving involuntary body ftuctuations and the occasional perception of individual
head movements as disturbing external accelerations) can be explained by a transient decoupling of
efference and efference copy, leading to amismatch
between anticipated and actual motion (Brandt and
Dieterich 1986; Brandt 1997). Healthy people can
experience similar mild sensations of vertigo without simultaneous anxiety during astate of exhaustion, when the difference between voluntary head
movements and involuntary ftuctuations become
blurred. In phobie patients this partial decoupling
(Fig. 32.3) may be caused by their preoccupation
with constant anxious controlling and retrospective
checking of balance. This leads to the perception of
sensorimotor adjustments that would otherwise
occur unconsciously by means of learned and automatically released muscle activation programmes to
maintain upright posture.
The dream-like illusion that normal motorists

Vertigo

472

--->------------~----------

corollary
discharge

IRE-AFFERENCES

ITJ
DJ
DJ

efference

SELF-MOTION I

~
CENTRAL
STORE

Fig.32.3. Schematic diagram of the sensorimotor derangement, or the neural mismatch concept, of phobic postural vertigo as
caused by decoupling of the efference-copy signal. An active movement leads to sensory stimulation, the messages of wh ich are compared with a multisensory pattern of expectation calibrated by early experience of motions (central store). The pattern of expectation
is prepared by the efference-copy signal, wh ich is emitted parallel to and simultaneously with the motion impulse (efference) or by
vestibular excitation during passive vehicular transportation.lf concurrent sensory stimulation and the pattern of expectation are in
agreement, self-motion is perceived while "space constancy" is maintained.lf the efference-copy signal (corollary discharge) is inappropriate, there is a sensorimotor mismatch, and a self-generated head motion or body sway may be erroneously perceived as external perturbations, causing the subjective postural vertigo. (From Brandt 1996.)

Table 32.1.

Psychosocial and psychodynamic variables determining the


life situation at onset of phobie postural vertigo (42 patients)'
Women

Men

9
7

13

Psychologieal stressors
partnership
family
profession
social surroundings
serious somatic illness

20
1

Relation to life eyele phase


young adulthood
midlife
retirement

2
9

5
21
3

(onflict themes
c\oseness / distance
loss / mourning
separation / individuation
autonomy
self-esteem
oedipality

2
2
3

5
3
2
2

13
5

16

6
1

3
5
2

occasionally experience (the car is stationary and


the environment passes by as in a film on a screen,
i.e. the sudden transition of perception of selfmotion into perception of external object motion) is
compatible with this hypothesis. When the car is
travelling at a constant speed, the vestibular apparatus, which reacts only to acceleration, is not stimulated.
Then the perception of self-motion is exclusively
optokinetic, induced by the relative environmental
shift (linearvection, see p. 409). Linearvection is
probably caused by visual afferent modulation of the
structures that identify self-motion, usually activated
by the efference copy. A periodic decoupling would
lead to the loss of perception of self-motion. This
evokes uncertainty and anxiety in a patient driving a
car and the patient then feels like a driver in a visual
vehicle simulator. Some of the patients suffering
from PPV complain about such driving illusions,
which in part keep them from driving at high speeds
or on motorways.

Secondary gain
allowance / care
financial compensation
masked hostility
identity via sickness role

, More than one category for each patient is possible


From Kapfhammer et al. (1997).

Body sway in PPV


Postural sway during upright stance was analysed in
patients with PPV and in age-matched healthy volunteers (Krafczyk et al. 1998). Recordings were made
under different conditions (with the eyes open or
closed): when standing on a foam rubber pad with
the head upright, turned 20 right or left, or during

473

Phobie postural vertigo

1 Hz horizontal head oseillations. Sway analysis


included calculation of sway path, of root mean
square values of sway, and of the power spectrum of
sway in fore-aft and lateral directions. Ihere was a
significant increase in sway activity in the 3.5-8 Hz
frequency band in patients with PPV (Fig. 32.4).
Ihis, however, does not impair objective postural
stability. Increase in high er frequency sway activity
may simply refiect a change in postural strategy
rather than a sensorimotor dysfunction. Ihe
patient's conscious control of stance may augment
coactivation of anti-gravity muscles, a strategy
which is applied by normal subjects when performing dem an ding balancing tasks.

Table 32.2. Diagnostie classifieation of the syndrome of phobie postural


vertigo (PPV) aeeording to structured clinical interview (SCID) at onset of
condition
DSM III-R diagnoses (axis I)

Personality features (axis 11)

---------------

Major depressive episode


Dysthymic disorder
Generalised anxiety disorder
Panic disorder
With agoraphobia
(complete)
With agoraphobia
(partial)
PPV (monosymptomatic)
With agoraphobia

2
2
6
14
8

Compulsive
Narcissistie
Dependent
Histrionie
Avoidant
Passive-aggressive

20
7
3
2
6
4

4
18
15

From Kapfhammer et al. (1997).

PPV: A panie disorder?


It is tempting to label PPV a sub type of panic disorder in which vertigo is the major complaint
(Frommberger et al. 1995; Stahl and Soefje 1995). A
psychiatric evaluation of 42 patients (Kapfhammer
et al. 1997), however, revealed that PPV is a clinical
entity that should not be confused with panic disorder or agoraphobia. Diagnostic classification by
structured clinical interview techniques (SeID),
according to DSM-III-R criteria, revealed a high rate
of psychopathological co-morbidity in PPV (Iable
32.2). However, in the largest subgroup of patients
(43%) vertigo manifested without anxiety or panic
attacks. Ihus, PPV occurs with or without anxiety;

60
50
40
30

20
10
O +---~-1~-+---r~~--+-~F-~i~-1

3.5

4.0

4.5

5.0

5.5

6.0

6.5

7.0

7.5

Hz

8.0

Fig.32.4. Phobie postural vertigo (PPV): postural sway when


standing on a foam rubber pad with eyes closed. Average power
speetrum of the fore-aft body sway in patients (blaek line, top;
n = 12) and normal controls (grey line, bottom; n = 12) for the frequeney range between 3.5 and 8 Hz. Patients show inereased
aetivity in this higher frequeney range. (From Krafezyk et al. 1998.)

the key to diagnosis is the monosymptomatic subjective postural imbalance without falls, rather than
the presence of anxiety. Ihis makes necessary the
nosological differentiation from panic dis orders
with agoraphobia.
Ihe three described conditions that most resemble PPV are "space phobia" (Marks 1981), "Mal de
debarquement syndrome" (Brown and Baloh 1987;
Murphy 1993), and "visual vertigo" (Bronstein 1995;
Bronstein et al. 1997). Space phobia, a pseudoagoraphobic syndrome, describes a fear of absent
visuospatial support (open spaces) and of falling,
unlike the fear of public places found in agoraphobia.
Ihe average age of onset in space phobia is 55 years
(in agoraphobia, 24), and the condition shows a poor
response to psychotherapy. Ihis special form of late
onset phobia is often induced by an illness or fall
and is characterised by a constant uncertainty rather
than by vertigo attacks. Mal de debarquement syndrome refers to the sensation of swinging, swaying,
unsteadiness and disequilibrium that is commonly
experienced in sea travel immediatelyon disembarking and wh ich lasts a few hours (Gordon et al.
1995). It may persist for weeks to years in some individuals after returning to land (Brown and Baloh
1987; Murphy 1993). It is at times an incapacitating
condition, which is probably more common than
one would anticipate from the literature and affiicts
females more frequently (Mair 1996). Here the
period of motion exposure - causing aftereffects
(such as seamen's legs) - acts as the inducing stimulus for a persistent sensorimotor derangement, a
development that resembles PPV, but is initiated by
an unusual and unadapted vestibular stimulation.
Visual vertigo as described by Bronstein (1995) is a
heterogeneous syndrome with peripheral or central
aetiologies. It may occur if patients with balance disorder show high visual field dependence.
Ihere was a clear preponderance of obsessivecompulsive personality traits in the patient group

474

presenting with PPV, but also narcissistie, histrionie,


dependent, avoidant, and passive-aggressive traits
were present (Kapfhammer et al. 1997; Table 32.2).
Most of the patients with PPV seem to be highly
ambitious and have a pronounced super-ego (berIch), a tendency to set themselves high standards of
achievement, and strong motivation. They create the
impression of being open and communicative, even
aggressive, but also anxious, hesitant, and easily irritated by criticism. This personality structure reflects
the fact that the phobias lie midway between obsessional neuroses and hysterias. When the condition of
PPV becomes manifest, the patients seem to be
(reactively?) depressed and moody. Anxiety and
vertigo seem to torment them because normal motivation is retained, but the anxiety and avoidance
behaviour inhibit their activities. They get through
their daily duties, but without any pleasure or satisfaction; they often complain of affective lability and
libido disorder, but less frequently of sleep disorders
and fluctuations of mood in the course of the day,
which are generally experienced in endogenous
depression. Their single-minded preoccupation with
vertigo gives rise to surprisingly concrete ideas as to
the supposed underlying disease: usually a tumour, a
brain ischaemie disorder, or multiple sderosis.
In our study the condition had lasted an average
of 3 years before a definitive diagnosis was established. This often meant that the patients had had
many contacts with doctors of various specialities,
had undergone several detailed invasive diagnostie
procedures, and had received different diagnoses,
such as "cervieal vertigo" or "vertebrobasilar insufficiency:'
The psychiatrie diagnostic assessment according
to SCID (DSM II1-R) revealed that the neurological
syndrome of PPV on first admission to the dizziness
unit had to be dassified as a panie dis order in 14
(with agoraphobia in 12), a generalised anxiety
dis order in 6, and a depressive dis order in 4
(Kapfhammer et al. 1997). Eighteen patients suffered
from recurrent monosymptomatie attacks of vertigo.
Without exception, these sensations of vertigo were
experienced with feelings of great discomfort and
distress during the acute manifestation, were cognitively evaluated as signs of a serious illness, but were
not accompanied by other physieal symptoms, whieh
occur in panic states. None of these patients complained of excess anxiety or even panie during the
attacks. In this patient subgroup, we can rule out the
possibility of a "non-fearful type" of panie disorder
comprising the whole list of somatic symptoms
usually detectable in states of panie, with the exception of the subjective fear of catastrophe or death
(Beitman et al. 1987). With recurring attacks, however, 15 patients developed a phobic avoidance pattern,

Vertigo
which was finally dassified as agoraphobia. In one
case an acrophobie behaviour was additionally
recorded.
From the perspective of life cyde, the majority of
patients (men: 21; women: 9) had to deal with serious midlife problems (Table 32.1). The socioeconomic status of most patients was "married" or
"living in stable long-term relationships:' Both women
and men had been living in partnerships characterised by emotional doseness, family centredness,
and social stability when either negative life events
or normative developmental tasks of the life phase
disrupted the previous life style. They were typically
freelancers or employed (Kapfhammer et al. 1997).
As regards psychiatric family history, a much
higher prevalence of anxiety disorders should be
expected in first-degree relatives, if PPV is more
strongly connected with other main groups of anxiety
dis orders (Crowe 1990; Fyer et al. 1990). An additional variable differentiating PPV from other
groups of anxiety disorders is the more advanced
age at first manifestation. Whereas age of onset in
various forms of phobia or panic disorder is most
frequently in the third decade of life (Noyes and Holt
1994), the incidence of PPV seems to be highest in
the fourth and fifth decades (Huppert et al. 1994).
A psychosocial and psychodynamie assessment of
the patient's life situation at the onset of the condition showed that almost all patients of our follow-up
study had been confronted with serious crises that
had disrupted their previous way of life (Fig. 32.1).
There seemed to be a traditional distribution of
social role and psychosocial identity, with special
stressors in the field of partnership and family prevailing in women, on the one hand, and stressors of
professional performance in men, on the other
(Kapfhammer et al. 1997). We agree with EckhardtHenn et al. (1997) that PPV allows for a more subgroup-oriented psychiatric dassification. However,
there are three consecutive diagnostie steps:
1. Differentiate between psychogenie and organic

vertigo syndromes;
2. Differentiate between PPV and other psychogenic forms of vertigo and dizziness;
3. Differentiate various psychiatric abnormalities
whieh according to the DSM IV dassification are
associated with PPV (Brandt et al. 1997).

Thus, the entity PPV, if evaluated by DSM-III-R or


DSM-IV, can be assigned to various psychiatric categories. Although an association of PPV with anxiety
disorders is evident, not all patients present with
symptoms of anxiety or panie during attacks of vertigo. However, most patients develop a disabling
"phobie-avoidance pattern" with recurrent attacks.

Phobie postural vertigo

Differential diagnosis
Relevant differential diagnosis of PPV includes eentral vestibular dis orders (p. 167), hereditary or
aequired ataxias, senile gait dis orders (p. 386), bilateral vestibulopathy (p. 127), peripheral or eentral
vestibular paroxysmia (p. 117), basilar migraine
(p. 329), familial episodic ataxia (p. 365), perilymph
fistulas (p. 99), atypical Meniere's disease (p. 83), and
psychogenic vertigo other than PPV (p. 455).

Course and treatment


The average age of our patients with PPV was 41.5
years (male: 40.9; female: 42.0). Ages ranged from 18
to 70 years, with a clear peak for the fourth and fifth
decades in both males and females. Both sexes were
equally affected; there was a slight preponderance of
females (n = 81 vs n = 73) (Huppert et al. 1994; Fig.
32.1). The duration of the condition before definitive
diagnosis ranged from 1 week to 30 years, with a
mean of 38 months. The attacks recurred for months
or even years, sometimes in cycles. On the basis of
patient history and clinical examination, a considerable proportion of patients were diagnosed initially
to have an organic dis order of vestibular function
wh ich triggered the phobic condition (p. 471). The
constellation of psychogenic vertigo following
lesional vertigo is described in the literature (Pratt
and McKenzie 1958; Lilienfeld et al. 1988; Page and
Gresty 1985; Marks 1981; Huppert et al. 1995; see
p.457).
The first therapeutic session must follow the diagnostic work-up. For the patients to "believe" a psychogenic diagnosis, they must have a neurological
examination, vestibular testing, and possibly a brain
imaging. Our therapeutic regimen consists mainly in
providing them with a detailed explanation of the
mechanism that causes, and the factors that provoke,
PPV attacks, and thus relieving them of their fear of
a severe organic disease. We recommend a selfcontrolled "desensitisation" - within the context of
behavioural therapy - by repeated exposure to situations that evoke the vertigo. Such therapy is usually
achieved in two to three sessions. We do not subject
the patients to long-term psychotherapy. We attempt
to guide the patients by suggestion, assuring them
that the nature of the disease is known and that selfcontrolled therapy is possible. We dis courage the
patient's preoccupation with their symptoms
(decoupling of catastrophic thoughts), provided the
obsessional symptoms are not too severe. In cases of
severe obsession we administer antidepressant

475

agents such as amitriptyline, imipramine, or selective serotonin-reuptake inhibitors in addition to


psychotherapy. We advocate regular but not overly
strenuous physical activity to improve the sense of
diminished fitness.
We asked 78 patients with PPV and 17 patients
with psychogenic disorder of stance and gait (PSG)
to evaluate their condition 6 months to 5.5 years
after their original referral and short -term psychotherapy (Brandt et al. 1994). Two results seem
most important:
1. We had not misdiagnosed anyone with PPV (also

Kapfhammer's group did not have to revise the


initial diagnosis in their follow-up of 42 patients;
Kapfhammer et al. 1997).
2. The patients showed a favourable course with an
improvement rate of 72% (22% of patients
became symptom-free). The majority of patients
with PPV experienced complete remission or
considerable improvement, even if their condition
had las ted between 1-30 years prior to diagnosis
(Fig.32.5).
Age and sex did not influence the prognosis. The
course seemed to be independent of psychotherapy
or other therapies performed, and most of the
patients with the best outcome reported no therapy
following or after the suggestive short-term psychotherapy described above (Table 32.3). There was
a tendency toward better prognosis with shorter
duration of PPV. The improvement or remission
usually occurred within days to weeks following
diagnosis. Improvement usually remained stable and
ameliorated both professional and social activities
considerably. Whereas the majority of patients with
PPV feit fit for work, the majority of those with PSG
felt unfit for work (Table 32.4). Occasional relapses
were not analysed statistically because of the varying
lengths of the follow-up periods. In patients who
failed to respond initially, attempts with long-term
psychotherapies were ineffective. Thus, it seems
unlikely that the prognosis of PPV will be considerably
improved by long-term rather than short-term therapies. On the other hand, there are no controlled
studies available on the different effects of alternative therapies for PPV; thus, the success rate of therapy
must be compared with the frequently favourable
course of phobias left untreated (Ag ras et al. 1972;
Noyes et al. 1980). It is well-established that behavioural therapy (Barlow 1990) and antidepressants
(Ballenger et al. 1988; Lesser et al. 1988) are helpful
in the treatment of phobias.
Our first follow-up allows us to adopt a moderately
favourable view on the course of PPV, which might
be further improved by additional pharmacological

476

Vertigo

>
c..
c..

2
1

- ---

- : :- ... 1. - - ---.- --: ,- --- .1- :-.-.-- - - - --

1
(9
Cf)

c..

weeks

10

I J/11
12

months

I IJ11

10

20 30
years

--

- --

.- - - - -

Fig.32.5. Course of phobie postural vertigo (PPV) (top) and psyehogenie disorder of stanee and gait (PSG) (bottam) as a funetion of
the duration of the eondition before diagnosis and short-term treatment by suggestion (abscissa). Single dots represent individual
self-evaluations of 78 patients with PPV and 17 patients with PSG of their eondition at follow-up using a simple seale (1 = symptomfree; 2 = considerably improved; 3 = no ehange).ln general, prognosis seems to be inversely related to the duration of the disorder, but
in PPV, eomplete remission is possible even after 20 years. No patient with PSG was found symptom-free if the eondition had lasted
longer than 4 months. (From Brandt et al. 1994.)

Table32.3. Results of therapeutic measures during foliow-up of 78 patients with phobie postural vertigo (PPV)
and 17 patients with psychogenic disorders of stance and gait (PSG)'
Type oftherapy

PPV

PSG

Symptom-free

Improved

No change

Symptom-free Improved

No change

(n=7)

(n=39)

(n=22)

(n=3)

(n=9)

10
10
7

5
5
2
5

Psychotherapy
1
Pharmacological therapy 1
Physiotherapy
2
Alternative therapies
1
No therapy
11

3
13

(n=5)

7
4
4

'Some patients received more than one therapy, thus totals may exceed the number of patients.
From Brandt et al. (1994).

or behavioural therapies not yet proven in controlled


prospective studies. In their follow-up, Kapfhammer
et al. (1997) emphasise the significant psychological
problems at follow-up term (74%), despite the considerable rate of improvement in vertigo complaints
(79%; Table 32.5). Only 11 of 42 patients were free of
any coexistent psychiatrie symptoms and were in a

stable and balanced mood at follow-up. Anxious (n =


17) and depressive syndromes (n = 11) were reported to be the main psychological problems. On a
descriptive level of analysis, patients with predominantly dependent or avoidant personality traits
seemed to show a more negative course of illness
than patients with obsessive-compulsive features. A

Phobie postural vertigo

477
Table 32.4. Ability of 78 patients with phobie postural vertigo (PPV)
and 17 patients with psyehogenie disorder of stanee and gait (PSG) to
work during follow-up period

Fit for work


Partially fit for work
Unfit for work
Other (pensioners)

PPV (n= 78)

PSG (n= 17)

45

6
2
7
2

14
13
6

From Brandt et al. (1994).

Table 32.5. Classifieation of psyehopathologieal disability aeeording to DSM-III-R at follow-up with respeet to
the initial diagnostie group (eategory of course of illness as regards the syndrome of vertigo)
First admission

Follow-up

(eategory I to IV)

No psychiatrie disorder

(1,1)

--------------~-

Major depressive episode

Dysthymic disorder

Dysthymie disorder

(l1I,IV)

Generalised anxiety disorder

Generalised anxiety disorder

(l1,1I,11I,11I,1I1,IV)

Panie disorder

No psychiatrie disorder
Dysthymie disorder
Panie disorder +
Dysthmie disorder
Panie disorder (partial)
Agoraphobia

2
2

(11,111)
(l11,IV)
(l1I,11I,1I1,IV)

PPV (monosymptomatie)

No psychiatrie disorder
Dementia
Adjustment disorders with
anxiety (ehronie type)
Agoraphobia
Agoraphobia +
Dysthymie disorder

4
4

(11,11,111,111,)
(11,111)

(1,1,1,11,111,111,111)
(111)

4
7

(11,11,11,11)
(l1,IV,IV,IV)
(lV,IV)

Categories: I = symptom-free; 11 = long symptom-free intervals; 111 = moderate improvement; IV = ehronicpersistent vertigo.
From Kapfhammer et al. (1997).

pronouneed "somatic concept of illness" and persisting physical symptoms such as tinnitus, neckache,
and serious illness, which were not causally linked to
PPV but reinforced constant self-observation and
often triggered hyperchondriac ideas, significantly
contributed to a negative course of illness
(Kapfhammer et al. 1997).
A major difference between PPV and PSG in our
follow-up study (Brandt et al. 1994) was the influence of the duration of the condition on prognosis.
Complete remission of PSG was observed only if the
condition had been present less than 4 months; there
was no improvement if it had lasted longer than 2
years (Fig. 32.5). The literature on prognosis and

treatment of conversion disorders is somewhat confusing; there have been both favourable and pessimistic reports (Ford and Folks 1985; Lazare 1981).
There is a striking difference in the appearance of
PPV and PSG as regards how patients experience
and present their condition. PPV patients tend to
hide their problem and usually continue working,
although they suffer enormously from impending
failure. Patients with PSG exhibit their obvious disability, do not usually work, and seem to suffer less
(Brandt et al. 1994).
Table 32.6 summarises the information about PPV
given in this chapter.

478
Table 32.6.

Vertigo
Phobic postural vertigo

Definition
Spontaneous (sometimes stimulus induced) postural vertigo and unsteadiness in maintaining an upright position and in walking with (but also without)
excess anxiety in patients with obsessive-compulsive personality traits
Clinical syndrome
- Dizziness and subjective disturbance of balance while standing or walking, despite normal clinical balance tests
- Fluctuating unsteadiness in episodes or momentary perception of illusory body perturbations
- Although the attacks can occur spontaneously, there is usually a perceptual stimulus or social situation as provoking factors and a tendency for rapid
conditioning, generalisation, and avoidance behaviour to develop
- Anxiety during or after vertigo, but also vertigo attacks without anxiety
- Obsessive-compulsive type personality, labile affect, mild depression
- Onset of condition frequently follows organic vestibular disorders, psychosocial stressors, or illness
Incidence/age/sex
This second-most common cause of vertigo affects both sexes equally; there is a clear peak for the fourth and fifth decades in both males and females
Primary personality
Clear preponderance of obsessive-compulsive personality traits, but also narcissistic, histrionic, dependent, avoidant, and passive-aggressive traits are
present
Hypothetical pathomechanism
Anxious introspection induces a sensorimotor mismatch due to decoupling of efference and efference-copy signal. so that the patients experience active
head and body movements as passive perturbations or vertigo
Course and prognosis
Lasting years to decades if untreated and showing increasing avoidance behaviour
Management
Explanation of the psychogenic mechanism (relieving the patients of their fear of a severe organic disease) and short-term behavioural therapy (Uselfcontrolled desensitisation U) yield an improvement rate of about 70%
Differential diagnosis
- Central vestibular disorders
- Non-vestibular ataxia
- Bilateral vestibulopathy
- Vestibular paroxysmia
- Basilar migraine
- Perilymph fistula
- Atypical Meniere's disease
- Psychogenic vertigo other than PPV
- uVisual vertigo U

References
Agras WS, Chapin HN, Oliveau DC (1972) The natural history of
phobia.Arch Gen Psychiatry 26:315-317
Ballenger JC, Burrows GD, Du Pont RL (1988) Alprazolam in panic
dis order and agoraphobia: results from a multicentre trial. Arch
Gen Psychiatry 45:413-422
Barlow DH (1990) Long-term outcome for patients with panic disorder treated with cognitive-behavioral therapy. J Clin
Psychiatry 51 (SuppIA):17-23
Beitman BD, Basha I, FIaker G, De Rosear L, Mukerji V, Lamberti J
(1987) Non-fearful panic disorder: panic attacks without fear.
Behav Res Ther 25:487-492
Brandt Th (1996) Phobie postural vertigo. Neurology
49:1480-1481
Brandt Th, Daroff RB (1980) The multisensory physiological and
pathological vertigo syndromes.Ann NeuroI7:195-203
Brandt Th, Dieterich M (1986) Phobischer AttackenSchwankschwindel. Ein neues Syndrom. Mnch med
Wochensehr 128:247-250

Brandt T, Huppert D, Dieterich M (1994) Phobie postural vertigo:


a first follow-up. J NeuroI241:191-195
Brandt T, Kapfhammer HP, Dieterich M (1997) Phobischer
Schwankenschwindel. Eine weitere Differenzierung psychogener
Schwindelzustnde erscheint erforderlich. Nervenarzt 68:848-849
Bronstein AM (1995) Visual vertigo syndrome: clinical and posturography findings. J Neurol Neurosurg Psychiatry 59:472-476
Bronstein AM, Gresty MA, Luxon LM, Ron MA, Rudge P, Yardley L
(1997) Phobie postural vertigo. Neurology 49: 1480-1481
Brown JJ, Baloh RW (1987) Persistent mal de debarquement syndrome: a motion-induced subjective dis order of balance. J
OtolaryngoI8:219-223.
Crowe RR (1990) Panic disorder: genetic considerations. J
Psychiatr Res 24: 129-134
Eckhardt-Henn A, Hoffman SO, Tettenborn B, Thamlske C, Hopf
HC (1997) Phobischer Schwindel. Nervenarzt 68:806-812
Ford CV, Folks DGC (1985) Conversion dis orders: an overview.
Psychosomat 26:371-382
Frommberger U, Angenendt J, Berger M (1995) Die Behandlung
von Panikstrungen und Agoraphobien. Nervenarzt 66: 173-186
Furman JM, Jacob RG (1997) Psychiatrie dizziness. Neurology
48:1161-1166

Phobie postural vertigo


Fyer AJ, Mannuzza S, Gallops MS, Martin LY, Aaronson C, Gorman
JM, Liebowitz MR, Klein DF (1990) Familial transmission of
simple phobias and fears. Arch Gen Psychiatry 47:252-256
Gordon CR, Spitzer 0, Doweck I, Melamed Y, Shupak A (1995)
Clinical features of Mal de debarquement: Adaption and habituation to sea conditions. J Vestib Res 5:363-369
Gresty MA, Bronstein AM, Brandt Th, Dieterich M (1992)
Neurology of otolith function: Peripheral and central disorders.
Brain 155:647-673
Holst E v, Mittelstaedt H (1950) Das Reafferenzprinzip
(Wechselwirkung zwischen Zentralnervensystem und
Peripherie). Naturwissenschaften 37:464-476
Huppert D, Brandt T, Dieterich M, Strupp M (1994) Phobischer
Schwankschwindel. Nervenarzt 65:421-423
Huppert D, Kunihiro T, Brandt T (1995) Phobic postural vertigo
(154 patients): its association with vestibular disorders. J Audiol
Med 4:97-103
Kapfhammer HP, Mayer C, Hock U, Huppert D, Dieterich M,
Brandt T (1997) Course of illness in phobic postural vertigo.
Acta Neurol Scand 95:23-28
Krafczyk S, Schlamp V, Dieterich M, Haberhauer P, Brandt Th
(1998) Increased body sway at 3.5-8 Hz in patients with phobic
postural vertigo. Neurosci Lett 259:149-152
Lazare A (1981) Current concepts in psychiatry. Conversion
symptoms. N Engl J Med 305:745-748

479
Lesser IM, Rubin RT, Pecknold JC, Rifkin A, Sinson RP, Lydiard RB,
Burrows GD, Noyes R, Du Pont RL (1988) Secondary depression
in panic dis order and agoraphobia. I. Frequency, severity, and
response to treatment. Arch Gen Psychiatry 45:437 - 443
Lilienfeld SO, Jacob RG, Furman JMR (1988) Vestibular dysfunction followed by panic dis order with agoraphobia. J Nerv Ment
Dis 177:700-701
Mair IWS (1996) The mal de debarquement syndrome. J Audiol
Med 5:21-25
Marks IM (1981) Space "phobia": a pseudo-agoraphobie syndrome. J Neurol Neurosurg Psyehiatry 44:387-391
Murphy T (1993) Mal de debarquement syndrome: a forgotten
entity? Otolaryngol Head Neck Surg 109:10-13
Noyes R, Clancey J, Hoenk PR, Hymen DJ (1980) The prognosis of
anxiety neurosis. Areh Gen Psychiatry 37:42-50
Noyes R, Holt CS (1994) Anxiety dis orders. In: Winokur G,
Clayton PJ (eds) The medieal basis of psyehiatry, 2nd edn.
Saunders, Philadelphia
Page MGR, Gresty MA (1985) Motorist's vestibular disorientation
syndrome. J Neurol Neurosurg Psyehiatry 84:729-735
Pratt RTC, McKenzie W (1958) Anxiety states following vestibular
dis orders. Lancet II:347-349
Stahl SM, Soefje S (1995) Panie attaeks and panie disorder: The
great neurologie impostors. Semin NeuroI15:126-132

SECTION L
Physiological vertigo

Physiological vertigo syndromes occur in healthy


subjects due to unusual and therefore unadapted
stimulation of the multisensory system for static and
dynamic spatial orientations. They are induced by
intersensory or intrasensory mismatches. The most

prominent forms are height vertigo (p.418) and


motion sickness (p. 485), but physiological vertigo
syndromes also include visual (p. 409), somatosensory (p. 481), auditory, head-extension (p. 297), and
bending-over vertigo (p. 298).

483

Motion sickness

Motion sickness is induced during pa sive Iocomotion in vehicles.lt is generated either by unfamiliar
body accelerations, to which the person has therefore not adapted, or by an intersensory mismatch
involving conflicting vestibular and visual stimuli
(Dichgans and Brandt 1973, 1978; Benson 1977;
Reason 1978; Brandt and Daroff 1980; Crampton
1990). According to the "mismatch theory" (see also
p. 4), spatial orientation and perception of movement are di turbed by a conflict between stimuli,
when the muLtisensory motion signals do not correspond to the expected pattern of sensory signals
established from earlier experience with active locomotion. This may give rise to unpleasant illusions of
movement with consequences for posture and vehicle control (Dichgans and Brandt 1978; Leibowitz et
al. 1982) and result in motion sickness due to summation. This simple "sensory conflict" theory of
motion sickness has been questioned by those who
argue that there is no principled basis on which this
concept can distinguish between nauseogenic and
non-nauseogenic stimulus situations (Stoffregen and
Riccio 1991; Riccio and Stoffregen 1991).
Motion sickness is not a superfluous epiphenomenon of vestibular stimulation, but rat her - in a teleological sense - it is a meaningful warning signal to
withdraw the body from unusual stimulus situations
that cannot be integrated by adequate dynamic
spatial orientation. The first known reports on seasickness are found in the writings of Hippocrates.
The major symptom of seasickness, nausea, is
derived from the Greek word "naus" for ship. The
general term "motion sickness," coined by Irwin
(1881), stresses the phenomenological similarity of
air, sea, car and space sickness despite their different
modes of provocation.

fort, tiredness, periodic yawning, and pallor. An


increase in facial pallor is followed by cold sweat,
increased salivation, oversensitivity to smells, pains
in the back of the head, and feelings of pressure in
the upper abdomen. Finally, the central symptoms of
nausea, retching and vomiting develop with motor
incoordination, loss of drive and concentration,
apathy and fear of impending doom (Chinn and
Smith 1955; Tyler and Bard 1949; Money 1970).
Autonomie nervous system responses and motion
sickness symptom levels are significantly related
throughout the entire sickness (Stout et al. 1995).
Motion sickness is an acute dis order that spontaneously remits within hours to a day after the inducing motion ceases. Even when individuals report
subjective recovery from a nauseogenic motion challenge, there is still a prolonged underlying effect of
motion sickness that sensitises the response to subsequent motion for aperiod of several years
(Golding and Stott 1997). Graybiel and colleagues
(1968) used "minor" and "major" symptoms culminating in vomiting and retching to grade the severity
of motion sickness. Subjective estimates of wellbeing (Reason and Graybiel1970) or magnitude estimations of actual nausea, in relation to standard
stimulation (magnitude estimation according to
Stevens 1957), have been used to evaluate experimental motion sickness in humans (Brandt et al.
1971).
Motion sickness has been considered an evolutionary anomaly in that the mechanism normally
initiating vomiting protects the body or brain from
ingested poisons (Treisman 1977).

The clinical syndrome

Vomiting is usually the culminating event of a regular sequence of symptoms of severe motion sickness.
However, if there is motion stimulation during sleep,
vomiting may occur immediately after the individual
has awakened. It is weil known from space missions

The full picture of acute motion sickness develops


after initial symptoms of dizziness, physical discom-

Nausea and vomiting

485

486

and parabolic flight experiments that episodes of


emesis can occur without the usual accompanying
prodromal signs of motion sickness (Lackner and
Graybiel 1986a). Similarly, during severe forms of
central positional vertigo (p. 291), sudden vomiting
without well-developed warning symptoms can be
elicited by a single head tilt (Baker and Bernat 1985;
Arbusow et al. 1998). Postoperative nausea and vomiting are customary, particularly in females; there is
evidence that patients who identified movement as
provoking their postoperative nausea had a specific
susceptibility to motion siekness (Kamath et al.
1990; Watcha and White 1992). Morphine (Riding
1975) and other anaestheties may sensitise the
vestibular system - as a migraine attack does (see
p. 334) - so that movements that are normally tolerated without any discomfort cause motion sickness.
Vestibular autonomic regulation (p. 16) involves
cardiovascular and respiratory responses, swallowing, and vomiting. Anatomical and electrophysiological connections between the vestibular and
autonomic systems have been described.
Accordingly, electrical stimulation of the vestibular
nerve evokes bilateral responses of the following
nerves: recurrent laryngeal, superior laryngeal, pharyngeal branch of the vagus, glossopharyngeal, and
hypoglossal with a latency of about 15 ms (Siniaia
and Miller 1996). A possible function of vestibulorespiratory reflexes is to provide adjustments in
breathing and airway patency during movements
and changes in position, e.g. to counteract gravitational effects on the diaphragm by increasing intraabdominal pressure through abdominal muscle
activity (Yates and Miller 1996, 1998; Siniaia and
Miller 1996). A network of vestibuloautonomic projections has been identified in the brainstem and
cerebellum of animals. This network has three components: (1) vestibulo-autonomic regions in the
vestibular nuclei, (2) direct descending pathways to
the brainstem circuitry for vestibulo-autonomic
reflexes, and (3) ascending pathways to the
parabrachial nucleus (Balaban and Porter 1998).
Whereas the descending pathways regulate respiration and vomiting, the ascending pathways have
access to the hypothalamus, amygdala, and neocortex to modulate affective (anxiety disorder) and neuroendocrine vestibular responses. This convergence
of vestibular and autonomie activity is also consiste nt with the concept that visceral input
(Mittelstaedt 1996) contributes to the perception of
verticality (Balaban and Porter 1998).
Excitation of the vestibular nuclei and archicerebellar influences cause centrally induced vomiting,
the major symptom of motion sickness. The latter
probably have an inhibitory effect by non-specifically
activating the reticular formation, the chemo-

Vertigo
receptive trigger zone, the vomiting centre, and the
vagal nuclei. Borison and Wang (1949) reported that
the neurophysiological site of the vomiting centre is
in the parvicellular reticular formation of the
medulla oblongata. However, the precise neuronal
link between the vestibulocerebellum and the vomiting centre is still missing (Mehier 1983; Yates and
Miller 1996).An intact chemoreceptive trigger zone
in the lateral postremal area (dorsolateral of the
IVth ventricle and the vagal nuclei) and the integrative vomiting centre (Cummings 1958), located in
the reticular formation alongside the solitary fascicle,
are suspected to playan important role.
Animal studies support the current view that
emesis is coordinated not by a unique, well-defined
vomiting centre, but rather by a distributed control
system located in the medulla between the levels of
the obex and retrofacial nucleus (Miller et al. 1994).
Neural activity in the brainstem midline region
(Miller et al. 1996), in the pre-Btzinger complex
(Zheng et al. 1997; Nakazawa et al. 1997) and in the
descending projections from the medullary nucleus
retroambigualis (Umezaki et al. 1997) possibly facilitates the input for the vomiting process.
Different vestibular brainstem pathways wh ich
subserve various functions receive projections from
different cerebellar regions: the vestibulo-ocular
reflex from the flocculonodular lobe, and the
vestibulospinal reflex from the anterior cerebellar
lobe (Balaban and Porter 1998). It is quite likely that
vestibulo-autonomie pathways also receive separate
archicerebellar projections. Four medial cerebellar
regions appear to be involved: (1) the lateral
nodulus-uvula, (2) the medial uvula, (3) the rostral
posterior lobe, and (4) the medial anterior lobe
(Ba1aban and Porter 1998). The obvious dissociation
of vomiting and nystagmus in single patients with
central vestibular positioning vomiting (p. 293) supports the view of functionally separate archieerebellar projections to the vestibulo-ocular reflex and the
vestibulo-autonomic reflex of vomiting. That positioning vomiting but not positioning nystagmus may
respond in these rare patients (Arbusow et al. 1998)
to benzodiazepines may be due to the different
neurotransmitters involved. Histamine Hl-receptors
stimulate the "emetic process" independently of
dopamine Dz-receptors in the chemoreceptive trigger zone and serotonin 5-HT 3-receptors in the visceral afferents, which are also involved in emetic
reflexes (Takeda et al. 1993). Antihistamines block
emetic Hl-receptors and thereby prevent motion
sickness.
The emetic response of an im als and humans to
certain poisons is also related to vestibular function.
Since the blood-brain barrier is less patent in the
postremal area, the chemoreceptive trigger zone is

487

Motion sickness

readily accessible to emetic substances. The severity


of the response to apomorphine differs according to
the position of the head in relation to the gravitational vector (Isaacs 1957). When the vestibular
apparatus of the inner ear is surgically removed
from dogs, the emetic response to lobeline,
levodopa, and nicotine is impaired (Money and
Cheung 1983). However, labyrinthectomy in dogs did
not affect pilocarpine- and apomorphine-induced
vomiting (Wang and Chinn 1956; Money and
Cheung 1983).

Labyrinth function and motion


sickness
Following the discovery in the nineteenth century of
the mechanism and functional importance of the
labyrinth for the perception of motion and the
maintenance of balance (Goltz 1870; Crum Brown
1874; Mach 1875), James (1882) realised from observations of deaf mutes that motion sickness cannot
occur unless the labyrinth is in functioning order.
This was later confirmed for patients with labyrinth
defects (Graybiel1964; Colehour 1965; Kennedy et
al. 1968). Animal experiments also demonstrated the
possibility of permanent insensitivity to motion
sickness. Both dogs and monkeys no longer had
motion sickness following bilateral sectioning of the
eighth nerve (Tyler and Bard 1949) or total
labyrinthectomy (Sjberg 1931; McNally et al. 1942;
Babkin and Bornstein 1943; Wang and Chinn 1956;
Kurashvili 1963; Johnson et al. 1962). The functional
integrity of both the semicircular canals (Miller and
GraybieI1972) and the otoliths is significant for the
pathogenesis of motion sickness. Resistance to
motion sickness was achieved in dogs by isolated
elimination of the six semicircular canals, but it
must be questioned whether the otoliths were really
spared by this method and were able to function
properly (Money and Friedberg 1964). Furthermore,
previously motion-emetic-sensitive squirrel monkeys
were rendered refractory to a standard motionemetic regimen by a two-stage utriculosacculectomy
that preserved the cristae ampullares of the semicircular canals (Brizzee and Igarashi 1986).
In comparison with the labyrinthine function,
"vision" is only a secondary rather than an essential
aetiological factor in the genesis of motion sickness.
This was indicated by Desnoes's observation (1926)
that it is quite possible for the blind to suffer motion
sickness and by Graybiel's comparative studies
(1970) which showed that the susceptibility of the
blind to seasickness did not significantly differ from

that of the non-blind. Nevertheless, there is a purely


"optokinetic motion sickness" (see Visual vertigo,
p. 409), which can be initiated in objectively stationary people. It is caused by a visual-vestibular convergence that has been demonstrated psychophysically
and in experiments with microelectrodes (Dichgans
and Brandt 1978).
For simple (linear and rotary) body oscillations,
the incidence of motion sickness decreases as the
frequency of the oscillation increases (Benson 1973).
A low frequency of about 0.2 Hz is most likely to
cause the syndrome, particularly for vertical motion.
Horizontal motion with subjects seated upright is
also nauseogenic, but the relationship of frequency
to nauseogenicity is less steep than that reported for
vertical motion (Golding and Markey 1996). The
most provocative situations arise from crosscoupled accelerations when the head, while undergoing a rotation about one axis, is bent about a second
axis; the extreme spatial disorientation that results is
called the "Coriolis effect" (p. 489) (Schubert 1931).
It is of particular concern to airplane pilots. Nausea
evoked by Coriolis stimuli is principally caused by
gyroscopic angular accelerations (Isu et al. 1994).
While the precise location of the body at the centre
of rotation is not critical during Coriolis stimulation,
the direction of head movement has a great effect on
nausea (Woodman and Griffin 1997). Passive rotation of the body about an off-vertical or horizontal
axis (barbecue-spit rotation) induces illusory perceptions ofbody movement and is a powerful stimulus of motion sickness (Guedry 1965; Leger et al.
1981; Denise et al. 1996). The sensory confiict that
elicits motion sickness may be visual-vestibular or
intravestibular (semicircular canals versus otoliths).
Even differences in otoliths and abdominal visceral
graviceptor dynamics have been accused of causing
sickness and erroneous perceptions of body orientation during exceptional transportation, e.g. helicopters and parabolic flight manouevres (Gierke and
Parker 1994). Vibration, if applied to the head, can
cause sensory illusions and motion sickness
(Lackner and GraybieI1974).

The sensory conflict theory (visualvestibular mismatch)


Although optokinetic stimuli are not essential conditions for the genesis of motion sickness, they may be
the sole cause and the decisive factors determining
the extent of the motion sickness when there is
simultaneous visual, vestibular, and somatosensory
stimulation. This can be seen from Coriolis

488

experiments in a revolving chair-revolving drum


system (Brandt et al. 1971; Dichgans and Brandt
1973) and from road-test results (Probst et al. 1982).
As a result of the initial observations by Armstrong
(1939), Manning and Steward (1949), Johnson and
Taylor (1961), Brown and Crampton (1966), and
Marshall and Brown (1966) on the inhibition of
motion sickness by visual motion control, it has
been suggested that people suffering from seasickness should be brought onto the deck so that they
can observe the motion of the ship relative to the
horizon (HillI937; Tyler 1946; Bruner 1955). There
has even been discussion about the usefulness of an
artificial horizon for submarines. Based on experimental testing of visual-vestibular interaction under
laboratory conditions (Brandt et al. 1971; Dichgans
and Brandt 1973; Lackner and GraybielI979), these
findings have led to the general recommendation of
visual control of motion during transportation.
The dependence of the intensity of motion sickness on the simultaneous visual stimulus conditions
(determined in road tests during linear accelerations
in cars; Probst et al. 1982) can be best explained by
an intersensory conflict of stimuli (Dichgans and
Brandt 1978; Oman 1982) or mismatch (Benson
1977; Reason 1978). As shown in the diagram (Fig.
1.2 in the Introduction, p. 5), active movement of the
body leads to stimulation of the three main
stabilising systems of the sens es (vestibular, visual,
somatosensory). The pattern of these sensory inputs
is compared in the central nervous system with a
multisensory pattern of expectation, previously calibrated for the motion. This pattern is evidently
retrieved from storage in the case of active body
movement by means of an efference copy emitted in
parallel with the initiation of the movement (von
Holst and Mittelstaedt 1950). If the actual sensory
inputs agree with the pattern of expectation, the perception of motion is controlled while the spatial constancy is maintained. However, if there is an
incongruity, e.g. between visual and vestibular information, there is a loss of spatial constancy, and a
feeling of disorientation and vertigo arises. The
summation of individual disorientation stimuli can
lead to the full picture of motion sickness.
When there are repeated, identical, incongruent
sensory inputs (e.g. Coriolis vestibular training of
pilots), the stored pattern of expectation is altered
for this type of movement. This explains the habituation of sailors and pilots to specific acceleration
stimuli. However, nausea and motion sickness are
not essential conditions for a plastic recalibration, as
Melvill Jones and Mandl (1980) demonstrated using
inversion prism glasses. When inversion prisms are
worn, motion sickness initially occurs because of the
incongruent retinal image migration as a result of

Vertigo
active head movements. This, however, does not
happen under stroboscopic ambient light: under
stroboscopic visual stimulus conditions, the vestibuloocular reflex adapts to the revers al of the direction
of the retinal image shifting. Takahashi and coworkers (1991, 1994, 1997) found that horizontal
reversion of vision resulted more often in motion
sickness than did vertical reversion. The authors
related this finding to the less disturbing consequences of the latter condition for postural control
and spatial orientation.
Cross-coupled accelerations acting on the
labyrinth cause apparent tilting sensations and nausea if the head is moved out of the axis of rotation
during uniform rotation (Coriolis effect, p.489).
Laboratory experiments using a combined revolving
chair-revolving drum system have shown that the
intensity of Coriolis effects brought about by
vestibular rotational accelerations is significantly
modulated by the simultaneous visual stimulus
(Brandt et al. 1971; Dichgans and Brandt 1973).
Dizziness and nausea were least pronounced in the
case of seat rotations, when there was adequate visual
perception of movement (in the light). They were
most pronounced in a coupled seat and drum rotation situation, when the visual signals of apparently
absent motion conflicted with the strong labyrinth
acceleration stimuli (Fig. 33.1).
As shown diagrammatically in Figs 33.1 and 33.2,
the multisensory pattern of expectation is retrieved
through the vestibular acceleration message during
passive body acceleration in a car. This is analogous
to the efference copy in active movement. When the
eyes are open, the visual movement message corresponds to the vestibular and somatosensory afferences, which are input at the same time. During the
"eyes-closed" condition, the redundancy of the
movement-signalling channels is reduced to two
(vestibular and somatosensory). Severe motion sickness with "stationary visual contrasts" in the visual
field of view results from the incongruity between
the expected visual stimuli compared via the initial
vestibular motion message (of a relative shifting of
the surroundings in the opposite direction) and the
signals of stationary surroundings being input at the
same time.
Squirrel monkeys also suffer from motion sickness, when combined vestibular and visual (optokinetic) stimuli are given in a direction and phase
mismatching mode (Igarashi et al. 1983). The "conflict sickness symptom score" is significantly higher
in the pitch plane than in the yaw plane in squirrel
monkeys (Igarashi et al. 1986a). Also in humans,
head movements during vection in pitch were most
stressful for motion sickness, followed by rollvection, then circularvection; however, circularvection

Motion sickness

489

Motion sickness
(angular oscil lation
of the body)

Coriolis effects
(head tilt in roll
du ring body rotation)

Magnitude estimation

c=

Apparent lilt
c::::::J Nausea

10 15 20

Motion sickness (magnitude estimation)

Fig. 33.1. Visual influences du ring body acceleration in a combined drum and chair system and their effects on vertigo and motion
sickness. Left. Magnitude estimations of apparent tilt and nausea induced by Coriolis effects (sideward movement of the head during
simultaneous chair and visual surround motion) at constant angular velocity of 60/5. Right. Magnitude estimation of motion sickness
induced by 15 min of sinusoidal angular oscillation of the body at 0.02 Hz and peak velocity of 100/5. The three visual conditions were
eyes open (top), rotary chair and visual surroundings mechanically coupled (middle), and eyes open in complete darkness (bottam).
Estimations of nausea indicate that the symptom is maximal with conflicting visual-vestibular stimuli (combined movement of chair
and drum) when vestibular acceleration disagrees with the visual information of no movement (Brandt 1976).

in yaw elicited the strongest sensation of selfrotation (Tiande and Jingshen 1991). The incidence
of motion sickness is also higher for true angular
oscillation in pitch than in yaw. This can be attributed to the more complex sensory mismatch that
occurs when semicircular canals and otoliths are
involved.
In the case of optokinetic vehicle simulator
motion sickness (McCauley 1984), the sensory pattern of expectation is retrieved optokinetically
through a visually induced, apparently spontaneous
movement (circularvection, linearveetion). The
absent vestibular and somatosensory aeeeieration
stimuli lead to an intersensory eonfliet. Motion siekness in high-fidelity visual simulators is a byproduet
of modern simulation teehnology, including virtual
environment systems (Kennedy et al. 1992, 1993).
Ski siekness (Husler 1995) has been reported as
a curiosity. It may manifest during down-hill skiing
on foggy days when visibility is redueed and thus
eontradietory multisensory information is provided.

Vestibular hyperexcitability
Transient hyperexcitability of the vestibular system
causes hypersensivity to head accelerations. Normal
active or passive head movements during locomotion may then lead to unusual and therefore
unadapted sensory responses with subsequent
motion siekness. Here the inherent properties of the
sensory system rather than the charaeteristics of the
motion stimulus are causative. This situation also
ereates a perceptual and sensorimotor eonfliet
beeause of the mismateh between the expeeted and
the aetually transdueed vestibular input. The gain of
the vestibulo-ocular reflex was found to be higher in
subjeets suseeptible to seasiekness (Gordon et
al. 1996), and a positive eorrelation was found
between suseeptibility to features of the Mal de debarquement syndrome (p.473) and seasiekness
(Gordon et al. 1995). Examples for transient vestibular
hyperexcitability-indueed motion siekness due to
physiologieal motion stimulation are migraine

Vertigo

490

g(mls'j

12
0.8

ACCEl..EAATION

,<x . dlrec:llon)

04

o O +-----:--L~,-;O;-,-t
04
0.8
1.2

ey open

oyes elosed

station.ary visu.. 1
surrounding5

Fig. 33.2. Visual influences on motion sickness (magnitude


estimation) induced in automobile passengers by repetitive
braking manoeuvres (linear accelerations along the horizontal
x-axis: 0.8 to 1.2 g) under real road conditions. The symptom is
maximal with conflicting visual-vestibular stimuli. Thus, providing ample peripheral vision of the relatively moving surroundings is the best way to physically alleviate car sickness. (From
Probst et al. 1982.)

attacks (p. 334) and postoperative nausea and vomiting. In a teleological sense, anaesthetics-induced
motion sickness supports Treisman's theory (1977)
that there are four different kinds of defence against
poisons in the stomaeh: (1) rejection by taste, (2)
vomiting provoked by effects on the stomach lining,
(3) vomiting provoked by stimulation of appropriate
chemoreceptors, and (4) vomiting provoked by a
mechanism such as motion sickness (Money 1990).

Incidence and susceptibility


Every individual can be made motion-siek, although
there is considerable interindividual variation to
susceptibility. The incidence of motion sickness at
sea varies from less than 1% to almost 100% (Money
1970). A questionnaire-based survey of motion sickness on board passenger ferries in the seas around

Great Britain revealed an average occurrence of


vomiting of 7% (Lawther and Griffin 1988). On an
Atlantic crossing with moderate turbulence, about
25% of the passengers experience motion sickness
within the first 3 days, whereas seasickness among
aircrew who survive in life rafts is reported to be as
high as 60% (Llano 1955; Money 1970).
It is weil known that susceptibility is greater in
females than in males and that the incidence of
motion sickness decreases with increasing age. The
question of possible connections between motion
sickness susceptibility, personality, and autonomie
function has been repeatedly raised (Gordon et al.
1994). Infants below the age of 2 years are highly
resistant to motion sickness because they use the
visual system only to a limited extent for dynamic
spatial orientation and are, therefore, subject to less
visual-vestibular perception conflicts (Brandt et al.
1976). Children older than 2 years are more susceptible than adults; this sensitivity peaks at around age
10 to 12 years. The incidence of motion sickness was
found to be an associated feature in 45% of cases of
childhood migraine, in contrast to an incidence of
5-7% in the control group (Barabas et al. 1983).
Motion sickness-like symptoms are part of migraine
attacks and probably caused by hyperexcitability of
vestibular (and other sensory) receptors secondary
to neuropeptide release (see Chap. 20, p. 334).
Children with Tourette's syndrome were found to
have a particularly high susceptibility to motion
sickness, which was hypothesised to be due to
neurotransmitter disturbances (Barabas et al. 1984).
This finding is, however, more probably explained by
the increased amount of head movements in these
patients, which result in cross-coupled accelerations
during passive transportation. Finally, individuals
with high aerobic fitness appeared to have an
increased susceptibility to motion sickness induced
by Coriolis (cross-coupled) vestibular stimulation
(Banta et al. 1987) . The heightened response of this
group to sensory conflicts may be related to particular patterns of sensory expectation (p. 5) which were
programmed by the extreme training. This agrees
with earlier findings that experienced pilots are
more likely to succumb to simulator sickness
(p. 417) than are trainee airmen.
Other species also exhibit motion sickness: dogs
have about the same susceptibility as humans.
Monkeys, horses, cows, seals, sheep, cats, birds, and
even fish have been reported to develop motion sickness and vomiting, whereas rabbits and guinea-pigs
appear highly resistant or insensitive (Money 1970).
Documented differences in stimuli that provoke
motion sickness exist even between various monkey
species (Corcoran et al. 1990).

Motion sickness

Management: physical and medical


prevention
Physical prevention
In the long term, physieal prevention is the most
effective therapeutie measure, if it involves vestibular training to promote central habituation (see Fig.
l.2, p. 5). During the actual situation of stimulation,
avoidance of head movements or use of head supports during acceleration in vehicles helps to reduce
the severity of motion siekness, because they help
avoid additional complex stimulations (e.g. crosscoupled accelerations), particularly in the case of
rotational accelerations when travelling around
bends. Vestibulospinal reflexes and learned posture
regulation, whieh counterregulate the body oscillations induced by vehicle accelerations, mayaiso help
(Fukuda 1975). The way in whieh the head is held
relative to the gravitation vector and to the direction
of acceleration is also important. A different habituation is apparently acquired during the stimulation of
movement along specific head axes (vertical, up or
down, horizontal transportation) under natural conditions. There is apparently a greater resistance to
stimuli along the z-axis. Susceptible individuals
should therefore be transported in helicopters and
rescue aircraft in a seated position with the head
held upright (von Baumgarten et al. 1980). On the
other hand, when such people are transported on
ships and in cars, experience shows that a recumbent
position is more favourable. Moreover, the head of
the patient being transported in an ambulance
should point in the direction of locomotion.

Visual prevention of motion sickness in vehicles


Road car tests with predominantly linear accelerations have shown that motion siekness can be signifieantly reduced by adequate visual control of the car
and body accelerations in contrast to the visual stimulus condition "eyes-closed" (Fig. 33.2). On the other
hand, the susceptibility to and severity of motion
siekness in cars are significantly increased if the field
of view is filled with largely stationary vehicle contrasts. This agrees with the experience made during
car journeys that the passenger in the back seat, particularly when reading, is at a greater risk of developing motion siekness than the driver who perceives
the relative movement of the surroundings. These
results confirm earlier recommendations for visually
preventing motion siekness, which were based on
laboratory experiments on the visual-vestibular
interaction in Coriolis effects and body oscillations

491

(Brandt et al. 1971; Diehgans and Brandt 1973;


Brandt and Daroff 1980). We consider it necessary to
extend these laboratory studies by carrying out road
tests to show that the visual modulation of vestibular
motion siekness is applicable not only to angular
accelerations but also to predominantly linear accelerations, which for example predominantly stimulate
the otoliths. Earlier laboratory investigations also
showed that dynamic, visual spatial orientation and
the optokinetie induction of sensation of one's own
movement depend, to a critical extent, on the magnitude of the stimulus field and particularly on the
stimulation of the peripheral field of view (Brandt et
al. 1973; Diehgans and Brandt 1978). It is also important that sufficiendy large and partieularly peripheral
parts of the field of view are available for visual control of vehicle movements. In vehicles or under
conditions in whieh visual perception of acceleration
is not possible, it is recommended that the eyes be
closed. In this way, it is possible to prevent motion
siekness in vehicles or at least to reduce its severity
by purely visual means.
Some researchers reported that autogeniefeedback training to control autonomie responses
under "Coriolis stimulations" reduced the incidence
of motion siekness to some extent (Toscano and
Cowings 1982; Cowings and Toscano 1982).
Subsequent studies have not confirmed this finding
(Jozsvai and Pigeau 1996).

Medical prevention
Anti-motion-siekness drugs such as scopolamine
(Transderm-Scop) and dimenhydrinate (Dramamine)
effectively prevent both vestibular and optokinetie
motion siekness (Brandt et al. 1974). Belladonna
alkaloids, such as scopolamine, were among the first
drugs to be used to prevent motion sickness and are
still among the most effective agents available. This
property of a rather selective suppression of motion
siekness was first discovered for a different pharmacologieal group of drugs, the antihistamines,
when using dimenhydrinate (Gay and Carliner
1949). Promethazine hydro chloride is the only
phenothiazine proven effective against motion siekness, and cinnarizine, a weak antihistamine with
some calcium antagonist properties, has the advantage of being well tolerated and without significant
side effects (Dobie and May 1994). Although cinnarizine is significantly superior to placebo in preventing seasickness (Doweck et al. 1994), it is obviously
less effective than scopolamine or dimenhydrinate.
The moderate anti-motion effect of ginger is probably due to its influence on the gastrie rather than the
vestibular system (Holtmann et al. 1989).

Vertigo

492

Dimenhydrinate in a single dose of 100 mg, and


scopolamine in a single dose of 0.6 mg or as
Transderm-Scop are among the most frequently
used anti-motion-sickness drugs. Whereas British
studies gene rally indicate a greater effectiveness of
the belladonna alkaloids, some US studies suggest
that the antihistamines are more potent (Wood et al.
1966a). Oral dosing of scopolamine shortly before
vomiting is largely useless, because gut motility is
suppressed when ill; however, apreparation of
scopolamine placed in the buccal pouch was shown
to be effective during parabolic flights (Norfleet et al.
1992). Since doubling the recommended dosages of
scopolamine failed to increase its efficacy (Wood et
al 1966b), it was tested in combinations with other
drugs for use in exceptionally severe motion sickness. Combinations with either D-amphetamine
sulphate and L-scopolamine hydrobromide (Wood et
al. 1966c; Wood and Graybiell968) or promethazine
with D-amphetamine provided far better protection
than any single drug (Wood and Graybiel 1970)
without increasing the side effects. Ephedrine did
not markedly increase the effectiveness of
scopolamine (Tokola et al. 1984). Anti-motionsickness drugs cause considerable central sedative
effects that measurably decrease performance
(Brandt et al. 1974; Wood et al. 1984; Parrot 1989).
Anti-motion-sickness drugs also impair ocular
motility, e.g. slowing of saccades (Brandt et al. 1974)
or visual fixation suppression (Collins et al. 1982). In
addition to nausea and vomiting, motion sickness
causes loss of performance, slowing of brain waves,
and inhibition of gastric motility. Ephedrine im (not
combined with scopolamine) is active against some
of these "secondary" effects of motion sickness
(Wood et al. 1990).
Table 33.1 summarises the information given
about motion sickness in this chapter.

Space sickness
Active head movements do not elicit motion sickness under gravitational conditions on earth but do
so in weightlessness: this is termed space sickness. In
the early Gemini and Mercury spacecraft programmes, astronauts were restrained in their seats
and did not develop symptoms, thus indicating that
a factor other than microgravity was involved. In
subsequent Apollo and Skylab missions, the astronauts were free to move. A dose relationship
between dizziness and active head movements was
observed (Graybiel et al. 1977; Graybiel 1979;
Lackner and Graybiel 1986b; Brandt and Daroff

Table 33.1.

Motion sickness

Syndrome
Initial symptoms such as dizziness, physical discomfort, tiredness,
yawning, facial pallor, cold sweat, occipital headache
Culminating in nausea and vomiting, apathy, and fear of impending doom
Incidence/age/sex
Every healthy individual can become motion siek, but there is a great
interindividual variability of susceptibility.
Susceptibility is greater in children and females than in adults and males.
Infants below the age oftwo are highly resistant.
Pathomechanism
Perceptual conflict or "mismatch" between actual and expected motion
stimulation
Intersensory (visual-vestibular) or intrasensory mismatch
Vestibular hyperexcitability (migraine attack, postoperative vomiting)
Course/prognosis
Spontaneous recovery after cessation of the provo king stimulation within
hours or a day
Spontaneous recovery also occurs with ongoing stimulation by central
habituation (readjustment of expectation to actual stimulation) within
three to six days
Management
Physical prevention
"Vestibular training" to promote central habituation avoidance of head
movements (Coriolis effects)
Alignment of head z-axis to preferred direction of acceleration
Adequate visual control of body motion (avoidance of visual-vestibular
conflict stimulation)
Medical prevention
Scopolamine (Transderm-Scop)
Dimenhydrinate (Dramamine)

1980). Some crewmembers showed areduction of


space sickness severity between the first and second
flights (Davis et al. 1988).
Two mechanisms have been proposed to explain
space sickness: sensory conflict and biological difference in the otoliths. In the sensory conflict
hypothesis, the symptoms are generated by an
"intravestibular" mismatch between the otoliths and
semicircular canals (Benson 1977) as well as
between otoliths and visual cues. The afferent consequences of self-generated body movements in space
are different from the signals expected from prior
calibration on earth. The semicircular canals and
otoliths correcdy transduce angular and linear accelerations, respectively, in space as on earth, but the
otoliths fail to signal orientation of the head in the
absence of gravity. Habituation resulting from recalibration of multisensory integration is established in
3-6 days and the symptoms due to head movements
abate for the remainder of the mission. However, the
return to earth causes transient vertigo that lasts
hours to days (Benson 1977; Graybiel et al. 1977).
Parabolic flight experiments have suggested a link
between the mechanisms evoking symptoms of

Motion sickness

space sickness and the mechanisms of nystagmus,


velo city storage and dumping (DiZio and Lackner
1991). Evidence supporting the sensory conflict
hypothesis for space motion sickness is provided by
experimental exposure of fishes to prolonged
weightlessness (Mori et al. 1996). Disturbance and
restoration of the dorsal light response (a functional
model of visual-graviceptor interaction) were seen
during the early phase of microgravity.
The second possible mechanism explaining space
sickness is related to the normal slight difference in
absolute weights of the two otoliths (von
Baumgarten and Thmler 1979). This biological difference is compensated for by central mechanisms in
the gravitational environment on earth. In the
microgravity of space, however, a recalibration is
required, and astronauts are symptomatic until it is
accomplished. Leigh and Daroff (1985) suggest that
careful clinical examination may be useful for
detecting those specific ocular motor abnormalities
that can be predicted by each hypothesis, such as
skew deviation, downbeating nystagmus, or
lateropulsion of saccades.
Finally, the patterns and levels of muscle innervation necessary to achieve particular body configurations and to bring about particular body movements
are greatly affected by background force level and
body orientation relative to the force vector
(Lackner et al. 1991). Altered sensorimotor control of
head and body posture may contribute to the manifestation of space sickness. Attempts have been made
to develop apparatuses and procedures to pre-adapt
astronauts to the sensory re arrangement associated
with weightlessness in space flight (Parker et al.
1987; Stern et al. 1989). Animal models established in
squirrel monkeys show that susceptibility to
vestibular-visual conflict sickness in pitch could be
effectively reduced by repeated exposure to the
provocative stimulation (Igarashi et al. 1986b).

References
Arbusow V, Strupp M, Brandt T (1998) Amiodarone-induced
severe, prolonged head-positional vertigo and vomiting.
Neurology: 51:917
Armstrong HG (1939) Principles and practice of aviation medieine. Williams and Wilkins, Baltimore
Babkin BP, Bornstein MB (1943) The effect of swinging and of
binaural galvanic stimulation on the motility of the stomach in
dogs. Rev Can Biol 2:336
Baker PCH, Bernat JL (1985) The neuroanatomy of vomiting in
man: association of projectile vomiting with a solitary metastasis in the lateral tegmentum of the pons and the middle cerebellar peduncie. J Neurol Neurosurg Psychiatry 48:1165-1168
Balaban CD, Porter JD (1998) Neuroanatomie substrates for
vestibulo-autonomic interactions. J Vestib Res 8:7-16

493
Banta GR, Ridley WC, McHugh J, Grissett JD, Guedry FE (1987)
Aerobic fitness and susceptibility to motion sickness. Aviat
Space Environ Med 58:105-108
Barabas G, Matthews WS, Ferrari M (1983) Childhood migraine
and motion sickness. Pediatrics 72:188-190
Barabas G, Matthews WS, Ferrari M (1984) Motion sickness in
children with Tourette's syndrome. Ann Neuro115:309
Baumgarten RJ von, Thmler R (1979) A model for vestibular
function in altered gravitational states. In: Holmquist R (ed)
(Cospar) Life sciences and space research. Pergamon Press,
Oxford, pp 161-170
Baumgarten RJ von, Baldrighi G, Vogel H, Thmler R (1980)
Physiological response to hyper- and hypogravity during
rollercoaster flight. Aviat Space Environ Med 51: 145-154
Benson AJ (1973) Physical characteristics of stimuli which induce
motion sickness, a review. IAM Rep 532:1-20
Benson AJ (1977) Possible mechanisms of motion and space sickness. Proceedings of the European symposium on life sciences
research in space. European Space Agency SP-130:101-108
Borison HL, Wang SC (1949) Functionallocalization of central
coordinating mechanism far emesis in cat. J Neurophysiol
12:305-313
Brandt Th (1976) Optisch-vestibulre Bewegungskrankheit,
Hhenschwindel und klinische Schwindelformen. Fortsehr
Med 94:1177-1182
Brandt Th, Daroff RB (1980) The multisensory physiologie al and
pathological vertigo syndromes.Ann NeuroI7:195-203
Brandt Th, Wist ER, Dichgans J (1971) Optisch induzierte PseudoCoriolis-Effekte und Circularvektion: Ein Beitrag zur optischvestibulren Interaktion. Arch Psychiat Nervenkr 214:365-389
Brandt Th, Dichgans J, Knig E (1973) Differential effects of central versus peripheral vision on egocentric and exocentric
motion perception. Exp Brain Res 16:476-491
Brandt Th, Dichgans J, Wagner W (1974) Drug effectiveness on
experimental optokinetic and vestibular motion sickness.
Aerospace Med 45:1291-1297
Brandt Th, Wenzel D, Dichgans J (1976) Die Entwicklung der
visuellen Stabilisation des aufrechten Standes beim Kind: Ein
Reifezeichen in der Kinderneurologie. Arch Psychiat Nervenkr
223: 1-13
Brizzee KR, Igarashi M (1986) Effect of macular ablation on frequency and latency of motion induced emesis in the squirrel
monkey. Aviat Space Environ Med 57: 1066-1070
Brown JH, Crampton GH (1966) Concomitant visual stimulation
does not alter habituation of nystagmic, oculogyral ar psychophysical responses to angular acceleration. Acta
Otolaryngol (Stockh) 61:80-91
Bruner JM (1955) Seasickness in a destroyer escort squadron. US
Armed Forces Med J 6:469-490
Chinn HI, Smith PK (1955) Motion sickness. Pharmacol Rev
7:33-83
Colehour JK (1965) Stress measurements in normal and
labyrinthine defective subjects in unusual force environments.
In: The role of vestibular organs in the exploration of space.
NASA, Sp-77, Washington DC, pp 347-355
Collins WE, Schroeder DJ, Elam GW (1982) A comparison of some
effects of three antimotion sickness drugs on nystagmic
responses to angular accelerations and to optokinetic stimuli.
Aviat Space Environ Med 53:1182-1189
Corcoran ML, Fox RA, Daunton NG (1990) The susceptibility of
rhesus monkeys to motion sickness. Aviat Space Environ Med
61:807-809
Cowings PS, Toscano WB (1982) The relationship of motion sickness susceptibility to learned autonomie control for symptom
suppression. Aviat Space Environ Med 53:570-575
Crampton GH (1990) Motion and space sickness. CRC Press, Boca
Raton
Crum Brown A (1874) On the sense ofrotation and the anatomy

494
and physiology of the semicircular canal of the internal ear. )
Anat PhysioI8:327-331
Cummings A) (1958) The physiology of symptoms: III Nausea and
vomiting.Am) Digest Dis 3:710-721
Davis )R, Vanderploeg )M, Santy PA, )ennings RT, Stewart DF
(1988) Space motion sickness during 24 flights of the space
shuttle. Aviat Space Environ Med 59:1185-1189
Denise P, Etard 0, Zupan L, Darlot C (1996) Motion sickness during off-vertical axis rotation: prediction by a model of sensory
interactions and correlation with other forms of motion sickness. Neurosci Lett 203:183-186
Desnoes Ph (1926) Seasickness. )AMA 86:319-324
Dichgans ), Brandt Th (1973) Optokinetic motion sickness and
pseudo-Coriolis-effects induced by moving visual stimuli. Acta
Otolaryngol (Stockh) 76:339-348
Dichgans ), Brandt Th (1978) Visual-vestibular interaction: Effects
on self-motion perception and postural contro!. In: Held R,
Leibowitz HW, Teuber HL (eds) Handbook of sensory physiology,
vol Vlll, Perception. Springer, Berlin Heidelberg New York, pp
755-804
DiZio P, Lackner JR (1991) Motion sickness susceptibility in parabolic flight and velocity storage activity. Aviat Space Environ
Med 62:300-307
Dobie TG, May)G (1994) Cognitive-behavioral management of
motion sickness. Aviat Space Environ Med (Suppl 10)
65:CI-C20
Doweck I, Gordon CR, Spitzer 0, Melamed Y, Shupak A (1994)
Effect of cinnarizine in the prevention of seasickness. Aviat
Space Environ Med 65:606-609
Fukuda T (1975) Postural behaviour in motion sickness. Acta
Otolaryngol (Stockh) 330:9-14
Gay LN, Carliner PE (1949) The prevention and treatment of
motion sickness. 1. Sea sickness. Bull lohns Hopkins Hosp
84:470-487
Gierke HE von, Parker DE (1994) Differences in otolith and
abdominal viscera graviceptor dynamics: implications for
motion sickness and perceived body position. Aviat Space
Environ Med 65:747-751
Golding JF, Markey HM (1996) Effect of frequency of horizontal
linear oscillation on motion sickness and somatogravic illusion.Aviat Space Environ Med 67:121-126
Golding )F, Stott )RR (1997) Objective and subjective time courses
of recovery from motion sickness assessed by repeated motion
challenges. ) Vestib Res 7:421-428
Goltz F (1870) ber die physiologische Bedeutung der
Bogengnge des Ohrlabyrinths. Pflgers Arch Ges Physiol
3:172-192
Gordon CR, Ben-Aryeh H, Spitzer 0, Doweck I, Gonen A,
Melamed Y, Shupak A (1994) Seasickness susceptibility, personality factors, and salivation. Aviat Space Environ Med
65:610-614
Gordon CR, Spitzer 0, Doweck I, Melamed Y, Shupak A (1995)
Clinical features of mal de debarquement: adaptation and
habituation to sea conditions.) Vestib Res 5:363-369
Gordon CR, Spitzer 0, Doweck I , Shupak A, Gadoth N (1996) The
vestibulo-ocular reflex and seasickness susceptibility. J Vestib
Res 6:229-233
Graybiel A (1964) Vestibular sickness and some of its implications
for space flight. In: Fields WS, Alfords BR (eds) Neurological
aspects of auditory and vestibular disorders. CC Thomas,
Springfield, III
Graybiel A (1970) Susceptibility to acute motion sickness in blind
persons. Aerospace Med 41 :650-653
Graybiel A (1979) Prevention and treatment of space sickness in
shuttle-orbiter missions. Aviat Space Environ Med 50: 171-176
Graybiel A, Miller EF, Homick )L (1977) Experiment M 131:
Human vestibular function. In: Biomedical results from skylab,
NASA Sp. 377, Washington DC, pp 74-103

Vertigo
Graybiel A, Wood CD, Miller EF, Cramer DB (1968) Diagnostic
criteria for grading the severity of acute motion sickness.
Aerospace Med 39:453-455
Guedry SE (1965) Orientation of the rotation-axis relative to
gravity: its influence on nystagmus and the sensation of rotation. Acta Otolaryngol (Stockh) 60:30-48
Husler R (1995) Ski sickness. Acta Otolaryngol (Stockh)
115:1-2
Hill J (1937) The care of the seasick. Br Med J:802-807
Holst E von, Mittelstaedt H (1950) Das Reafferenzprinzip
(Wechselwirkungen zwischen Zentralnervensystem und
Peripherie). Naturwissenschaften 37:464-476
Holtmann S, Clarke AH, Scherer H, Hhn M (1989) The antimotion sickness mechanism of ginger. Acta Otolaryngol
(Stockh) 108:168-174
Igarashi M, Isago H, O-Uchi T, Kulecz WB, Homick )L, Reschke MF
(1983) Vestibular-visual conflict sickness in the squirrel monkey.Acta Otolaryngol (Stockh) 95:193-198
Igarashi M, Kobayashi K, Kulecz WB, Isago H (1986a) Vestibularvisual conflict in pitch and yaw planes in the squirrel monkey.
Aviat Space Environ Med 57:1071-1074
Igarashi M, Kobayashi K, Kulecz WB, Himi T (1986b) Changes in
susceptibility to vestibular-visual conflict sickness in monkeys
by repeated exposure. Acta Otolaryngol (Stockh) 102:432-437
Irwin)A (1881) The pathology of seasickness. Lancet 11:907-909
Isaacs B (1957) The influence of head and body position on the
emetic action of apomorphine in man. Clin Sci 16:215-221
Isu N, Yanagihara MA, Mikuni T, Koo) (1994) Coriolis effects are
principally caused by gyroscopic angular acceleration. Aviat
Space Environ Med 65:627-631
)ames W (1882) The sense of dizziness in deaf-mutes. Am) Otol
4:239-254
Johnson WH, Taylor NBG (1961) The importance of head movements in studies involving stimulation of the organ of balance.
Acta Otolaryngol (Stockh) 53:211-218
)ohnson WH, Meek ) C, Graybiel A (1962) Effects of labyrinthectomy
on canal sickness in squirrel monkey. Ann Otol Rhinol
Laryngo171:289-298
)ozsvai EE, Pi ge au RA (1996) The effect of autogenic training and
biofeedback on motion sickness tolerance. Aviat Space Environ
Med 67:963-968
Kamath B, Curran J, Hawkey C, Beattie A, Gorbutt N, Guiblin H,
Kong A (1990) Anaesthesia, movement and emesis. Br) Anaesth
64:728-730
Kennedy RS, Graybiel A, McDonough RC, Beckwith FD (1968)
Symptomatology under storm conditions in the North Atlantic
in control subjects and in persons with bilateral labyrinth
defects. Acta Otolaryngol (Stockh) 66:533-540
Kennedy RS, Lane NE, Lilienthai MG Berbaum KS, Hettinger LJ
(1992) Profile analysis of simulator sickness symptoms: application to virtual environment systems. Presence 1:295-301
Kennedy RS, Lane NE, Berbaum KS, Lilienthai MG (1993)
Simulator sickness questionnaire: an enhanced method for
quantifying simulator sickness. Int J Aviat Psycho13:203-220
Kurashvili AE (1963) Vestibular reactivity during the cumulative
action of slow centripetal accelerations. Office of Technical
Services, FTD-MT-63-179
Lackner )R, Graybiel A (1974) Elicitation of vestibular side effects
by regional vibration of the head. Aerospace Med 45:1267-1272
Lackner )R, Graybiel A (1979) Some influences ofvision on susceptibility to motion sickness. Aviat Space Environ Med
50:1122-1125
Lackner )R, Graybiel A (1986a) Sudden emesis following parabolic
flight maneuvers: implications for space motion sickness. Aviat
Space Environ Med 57:343-347
Lackner )R, Graybiel A (1986b) Head movements in non -terrestrial
force environments elicit motion sickness: implications for the
etiology of space motion sickness. Aviat Space Environ Med
57:443-448

Motion sickness
Lackner JR, Graybiel A, DiZio P (1991) Altered sensorimotor control of the body as an etiological factor in space motion sickness.Aviat Space Environ Med 62:765-771
Lawther A, Griffin MJ (1988) A survey of the occurrence of
motion sickness amongst passengers at sea. Aviat Space
Environ Med 59:399-406
Leger A, Money KE, Landolt JP, Cheung BS, Rodden BE (1981)
Motion sickness caused by rotations about earth-horizontal
and earth-vertical axes. J Appl PhysioI50:469-477
Leibowitz HW, Post RB, Brandt Th, Dichgans J (1982) Implications
of recent developments in dynamic spatial orientation and
visual resolution for vehicle guidance. In: Wertheim AH,
Wagenaar WA, Leibowitz HW (eds) Tutorials on motion perception. Plenum Press, New York, pp 231-260
Leigh RJ, Daroff RB (1985) Space motion sickness: Etiological
hypotheses and a proposal for diagnostic clinical examina ti on.
Aviat Space Environ Med 56:469-473
Llano GA (1955) Airmen against the sea - an analysis of sea survival experiences. Maxwell AFB, Research Studies Institute,
ADTIC Publ G-104
Mach E (1875) Grundlinien der Lehre von den Bewegungsempfindungen. Engelmann, Leipzig
Manning GW, Steward WG (1949) Effect ofbody position on ineidence of motion sickness. J Appl Physioll:619-628
Marshall JE, Brown JH (1966) Visual arousal interaction and
specificity of nystagmic habituation. US Army Medical
Research Laboratory, Fort Knox, Report No 688
McCauley ME (1984) Research issues in simulator sickness.
Proceedings of a workshop. National Academy Press,
Washington, DC
McNally WJ, Stuart EA, Morton G (1942) Effect oflabyrinthectomy
on motion sickness in animals. In: Proceedings of the conference on motion sickness. National Research Council of Canada
Toronto, Report No. C-748
Mehler WR (1983) Observations on the connectivity of the parvicellular reticular formation with respect to a vomiting centre.
Brain Behav Evol 23:63-80
Melvill Jones G, Mandl G (1980) "Motion" sickness due to vision
reversal: its disappearance in stroboscopic light. Ann NY Acad
Sei 374:303-311
Miller AD, Nonaka S, Jakus J (1994) Brain areas essential or nonessential for emesis. Brain Res 647:255-264
Miller AD, Nonaka S, Jakus J, Yates BJ (1996) Modulation of vomiting by the medullary midline. Brain Res 737:51-58
Miller EF, Graybiel A (1972) Semieircular canals as a primary etiological factor in motion sickness.Aerospace Med 43:1065-1074
Mittelstaedt H (1996) Somatic graviception. Biol PsychoI42:53-74
Money KE (1970) Motion sickness. Physiol Rev 50:1-39
Money KE (1990) Motion sickness and evolution. In: Crampton
GH (ed) Motion and space sickness. CRC Press, Boca Raton, pp
1-7
Money KE, Friedberg J (1964) The role of the semicircular canals
in causation of motion sickness and nystagmus in the dog. Can
J Physiol PharmacoI42:793-801
Money KE, Cheung BS (1983) Another function of the inner ear:
faeilitation of the emetic response to poisons. Aviat Space
Environ Med 54:208-211
Mori S, Mitarai G, Takabayashi A, Usui S, Sakakibara M, Nagatomo
M, Baumgarten RJ von (1996) Evidence of sensory conflict and
recovery in carp exposed to prolonged weightlessness. Aviat
Space Environ Med 67:256-261
Nakazawa K, Zheng Y, Umezaki T, Miller AD (1997) Vestibular
inputs to bulbar respiratory interneurons in the cat. Neuro
Report 8:3395-3398
Norfleet WT, Degioanni JJ, Calkins DS, Reschke MF, Bungo MW,
Kutyna FA, Homick JL (1992) Treatment of motion sickness in
parabolic f1ight with buccal scopolamine. Aviat Space Environ
Med 63:46-51
Oman CM (1982) A heuristic mathematical model for the dynamics

495
of sensory conflict and motion sickness. Acta Otolaryngol
(Stockh) SuppI392:1-44
Parker DE, Reschke MF, Gierke HE von, Lessard CS (1987) Effects
of proposed preflight adaptation training on eye movements,
self-motion perception, and motion sickness, a progress report.
Aviat Space Environ Med 58:42-49
Parrot AC (1989) Transdermal scopolamine: a review of its effects
upon motion sickness, psychological performance, and physiological functioning. Aviat Space Environ Med 60: 1-9
Probst Th, Krafczyk S, Bchele W, Brandt Th (1982) Visuelle
Prvention der Bewegungskrankheit im Auto. Arch Psychiat
Nervenkr 231:409-421
Reason JT (1978) Motion sickness adaptation: a neural mismatch
model. J Roy Soc Med 71:819-829
Reason JT, Graybiel A (1970) Changes in subjective estimates of
well-being du ring the onset and remission of motion sickness
symptomatology in the slow rotation room. Aerospace Med
41:166-171
Riccio GE, Stoffregen TA (1991) An ecological theory of motion
sickness and postural instability. Ecol PsychoI3:195-240
Riding JE (1975) Minor complications of general anaesthesia. Br J
Anaesth 47:91-101
Schubert G (1931) ber die physiologischen Auswirkungen der
Coriolis- Krfte bei Trudelbewegungen des Flugzeuges. Acta
Otolaryngol (Stockh) 16:39-47
Siniaia MS, Miller AD (1996) Vestibular effects on the upper airway musculature. Brain Res 736:160-164
Sjberg AA (1931) Experimentelle Studien ber den
Auslsungsmechanismus der Seekrankheit. Acta Otolaryngol
(Stockh) Supp114:136
Stern RM, Hu S, Vasey MW, Koch KL (1989) Adaptation to
vection-induced symptoms of motion sickness. Aviat Space
Environ Med 60:566-572
Stevens SS (1957) On the psychophysical law. Psychol Rev
64:153-181
Stoffregen TA, Riccio GE (1991) An ecological critique of the
sensory conflict theory of motion sickness. Ecol Psychol
3:159-194
Stout CS, Toscano WB, Cowings PS (1995) Reliability of psychophysiological responses across multiple motion sickness
stimulation tests. J Vestib Res 5:25-33
Takahashi M, Ogata M, Miura M (1997) The significance of
motion sickness in the vestibular system. J Vestib Res 7:179-187
Takahashi M, Saito A, Okada Y, Takei Y, Tomizawa I, Uyama K,
Kanzaki J (1991) Locomotion and motion sickness during horizontally and vertically reversed vision. Aviat Space Environ
Med 62:136-140
Takahashi M, Toriyabe I, Takei Y, Kanzaki J (1994) Studyon experimental motion sickness in children. Acta Otolaryngol (Stockh)
114:231-237
Takeda N, Morita M, Hasegawa S, Horii A, Kubo T, Matsunaga T
(1993) Neuropharmacology of motion sickness and emesis.
Acta Otolaryngol (Stockh) SupplSOI:1O-J5
Tiande Y, Jingshen P (1991) Motion sickness severity under interaction of vection and head movements. Aviat Space Environ
Med 62:141-144
Tokola 0, Laitinen LA, Aho J, Gothoni G, Vapaatalo H (1984) Drug
treatment of motion sickness: scopolamine alone and combined with ephedrine in real and simulated situations. Aviat
Space Environ Med 55:636-641
Toscano WV, Cowings PS (1982) Redueing motion sickness: A
comparison of autogenic-feedback training and an alternative
cognitive task. Aviat Space Environ Med 53:449-453
Treisman, M (1977) Motion sickness: an evolutionary hypothesis.
Seience 197:493-495
Tyler DB (1946) The influence of placebo, body position, and
medication on motion sickness. Am J PhysioI146:458-466
Tyler DB, Bard P (1949) Motion sickness. Physiol Rev 29:311-369
Umezaki T, Zheng Y, Shiba K, Miller AD (1997) Role of nucleus

496
retroambigualis in respiratory reflexes evoked by superior
laryngeal and vestibular nerve afferents and in emesis. Brain
Res 769:347-356
Wang SC, Chinn HJ (1956) Experimental motion sickness in dogs.
Importance of labyrinth and vestibular cerebellum. Am J
PhysioI185:617-623
Watcha MF, White PF (1992) Postoperative nausea and vomiting.
Its etiology, treatment, and prevention. Anesthesiology
77:162-184
Wood CD, Graybiel A (1968) Evaluation of sixteen anti-motion
sickness drugs under controlled laboratory conditions.
Aerospace Med 39:1341-1344
Wood CD, Graybiel A (1970) Evaluation of anti-motion sickness
drugs: A new effective remedy revealed. Aerospace Med
41:932-933
Wood CD, Graybiel A, McDonough R (1966a) Human centrifuge
studies on the relative effectiveness of some anti-motion sickness drugs.Aerospace Med 37:187-190
Wood CD, Graybiel A, Kennedy RS (1966b) Comparison of effectiveness of some antimotion sickness drugs using recommended
and larger than recommended doses as tested in the slow rotation room. Aerospace Med 37:259-262

Vertigo
Wood CD, Kennedy RE, Graybiel A, Trumbull R, Wherry RJ
(1966c) Clinical effectiveness of anti-motion sickness drugs.
JAMA 198:1155-1158
Wood CD, Manno JE, Manno BR, Redetzki HM, Wood M, Vekovius
A (1984) Side effects of antimotion sickness drugs. Aviat Space
Environ Med 55:113-116
Wood CD, Stewart J], Wood MI, Manno JE, Manno BR, Mims ME
(1990) Therapeutic effects of antimotion sickness medications
on the secondary symptoms of motion sickness. Aviat Space
Environ Med 61:157-161
Woodman PD, Griffin MJ (1997) Effect of direction of head movement on motion sickness caused by Coriolis stimulation. Aviat
Space Environ Med 68:93-98
Yates BJ, Miller AD (1996) Vestibular respiratory regulation. In:
Miller AD, Bianchi AL, Bishop BP (eds) Neural control of the
respiratory muscles. CRC Press, Boca Raton, pp 271-282
Yates BJ, Miller AD (1998) Physiological evidence that the vestibular system participates in autonomic and respiratory contro!. J
Vestib Res 8:17-25
Zheng Y, Umezaki T, Nakazawa K, Miller AD (1997) Role of preinspiratory neurons in vestibular and laryngeal reflexes and in
swallowing and vomiting. Neurosci Let! 225:161-164

Index

Acetazolamide 49,91,372
Acetylsalicylate 328
ototoxicity 397,399
Acoustic neurinoma 159,160
Acrophobia 460-1
psychotherapy 461
Acute otitis media 147,149-51
Adaptation 55
vestibulo-ocular 60
Age-related changes
eye movements 385-6
postural sway and balance 386
vestibulo-ocular reflexes 385-6
Agoraphobia 18,459-60
criteria for 459-60
epidemiology 459-60
management 460
psychotherapy 461
Albers-Schnberg disease 145,148
Alcohol,ototoxicity 398,399
Alkylating agents, ototoxicity 397
Alport's syndrome 131
Alternobaric vertigo 351-2
Amitriptyline 328
otoxicity 398
Amphetamine,ototoxicity 393
Analgesics 328
Anterior inferior cerebellar artery
infarction 308-9
Anterior semicircular canal fistula 112
Antibiotics 49
ototoxic 49
Anticholinergic drugs 50
Anticonvulsants 49
Antiemetics 49-51,328
Antihistamines 50
Antivertiginous drugs 49-51
Anxiety neurosis 458
Approach to patient 23-48
auditory dysfunction 27
dizziness and light -headedness 23,24
dizziness and postural imbalance
27-8

neuro-ophthalmological and
otoneurological evaluation
34-46

orthostatic hypotension 25
oscillopsia 26-7
otolithic vertigo 29
paroxysmal vertigo 29-33
positional/positioning vertigo 26
recurrent vertigo attacks 25
semicircular canal vertigo and mixed
canal-otolith vertigo 28-9

sustained (rotatory) vertigo 24,26


syncope 25
Arnold-Chiari malformation 199,203
Arthrokinetic nystagmus 446-7
Ataxia
cervical vertigo 444
familial episodic 33, 173
familial periodic 365-74
paroxysmal 242
vestibular 4
Auditory dysfunction 27
Autoimmune inner ear disorders 131,
151-5

Bell's palsy 146


Bending-over vertigo 298
Benign paroxysmal positioning vertigo
14,15,33,251-87

aetiology 264-5
anterior semicircular canal BPPV
279-80

in childhood 335-6,376
clinical syndrome 252-4
positioning nystagmus 253-4
differential diagnosis 256-7
horizontal semicircular canal BPPV
269-79

aetiology and pathomechanism


271-8

Baclofen 49, 181


downbeat nystagmus 204, 205
Balance
age-related changes 380
extraocular muscle paresis and 429-30
Balance training
normal subjects 52-3
vestibular dis orders 53-5
Barbiturates,ototoxicity 397,398,399
Barotrauma 351-4
alternobaric vertigo 351-2
blast injury 348,352
decompression sickness 352
round/oval window fistula 353-4
Basilar artery infarction 312-14
Basilar migraine 327,329-33
age of onset 331-2
association with headache 332
association with other forms of
migraine 332
association with other vertigo dis orders

reversible ipsilateral caloric


hypoexcitability 275-8
transition of canalolithiasis to
cupulolithiasis 274-5
clinical syndrome 270-1
atypical h-BPPV with apogeotropic
positional nystagmus 270-1
management 278-9
natural course 271
iatrogenic 357
management 265-9
positional exercises and liberatory
manoeuvres 265-8
surgical procedures 269
natural course 256
pathomechanism 257-64
canalolithiasis 259-62
peripheral vs central vestibular
dysfunction 257
traditional view of cupulolithiasis
257-9

unilateral mimicking bilateral BPPV

337-8

benign paroxysmal torticollis in infancy


336

benign paroxysmal vertigo in


childhood 335-6
benign recurrent vertigo 337
central ocular motor signs 332
characteristics of 332
in childhood 370
clinical syndrome 335-8
diagnosis with episodic vertigo 332-3
dizziness and vertigo as facultative
symptoms 337
motion sickness-like symptoms 334-5
ocular motor deficits in symptom-free
interval 335
origin of vertigo in migraine 333-4
vestibular symptoms 332

497

262-4

post-traumatic 347
vertigo and posture 254-7
Benign paroxysmal torticollis, in in fants
336

Benign recurrent vertigo 337


Benzodiazepines 50
ototoxicity 398,399
Betahistine 49,91
Beta-receptor blockers 49,328
Bilateral vestibulopathy 14,15,127-45
aetiology and pathomechanisms
129-37

cerebellar degeneration 130


clinical syndrome 127-8
diagnosis 128-9,130
idiopathic 130, 132-5

498

Index

Bilateral vestibulopathy - cant.


immune-mediated 130
management 137-8
neoplastic 131,133
ototoxic 130
symptoms and differential diagnosis
129

Blast injury 348,352


Bodysway
flicker illumination 427
height vertigo 420
histogram 311
nystagmus 430
phobie postural vertigo 472-3
visual acuity 425,426

Borrelia burgdorferi

131

Brain stern auditory-evoked potentials 120


Brain tumours, and vestibular syndromes
244-5

Brandt-Daroff exercises 265


Bromides,ototoxicity 393
Buoyancy hypothesis 285-9
Calcium antagonists 328
Caloric testing 44
vestibular neuritis 69-70
Camptocormia 472
Camurati-Engelmann disease 123
Canalith jam 268
Canalith repositioning manoeuvre 267
Canalolithiasis 259-64
transition to cupulolithiasis 274
Carbamazepine 235
ototoxicity 397,398,399
paroxysmal vertigo 242
vestibular paroxysmia 122-3
Central brainstem/cerebellar lesions
241-2

Central positional nystagmus 292


Central positional vertigo 291-9
bending-over vertigo 298
central positional nystagmus 292
head (neck)-extension vertigo 297-8
paroxysmal positioning vomiting
293-6

positional downbeat nystagmus 291-2


transient vertebrobasilar ischaemia
296-7

Cerebellar intoxication 397-400


Cervical vertigo 441-6,473
ataxia and nystagmus in experimental
cervical vertigo 444
clinical evidence 444-6
neck afferents and neck reflexes 442-4
central pathways 443-4
cervico-ocular reflex 443
neck reflexes 443
spatial orientation 442-3
Cervico-ocular reflex 443
Channelopathies
calcium 370-2
potassium 370
Children
basilar migraine 370
benign paroxysmal torticollis in infancy
336

benign paroxysmal vertigo in


childhood 335
central vestibular syndromes 379

Meniere's disease 377


motion sickness 376
perilymph fistulas 377
peripheral vestibular failure 378-9
unilaterallbilateralloss of vestibular
function 377-8
vertigo in 375-81
vestibular neuritis 376
Chloral hydrate, ototoxicity 398
Cholesteatoma 147,151,152,153
Cholesterol cyst 154
Circularvection 409-16
arthrokinetic 446
psychophysics 411-13
rollvection-tilt 414-16
visual-vestibular interaction 413-14
Clonazepam 48
Cogan's syndrome 131,143,154-5,156,
157

Compensation 57
Computed tomography
labyrinth and vestibular nerve 143
perilymph fistula 102
Convergence-retraction nystagmus 4,37
Coriolis effect 487
Cortical astasia 192
Cortical infarctions 315-22
cortical rotational vertigo 318-22
Cortical rotation al vertigo 318-22
Corticosteroids 49,77
Cover tests 35
Cupulolithiasis 257-9,285-6
tradition al view 257-9
transient 262
transition from canalolithiasis 274
Dark, visual stabilization in 426
Decompression sickness 352
Delayed endolymphatic hydrops 88,348
Depression 458
Diazepam SO
Dihydroergotamine 328
Dimenhydrinate 50,76,491
otoxicity 398
Disabling positional vertigo see Vestibular
paroxysmia
Dix-Hallpike manoeuvre 42,252,253
Dizziness 23,24,27-8
in elderly 389-92
as facultative symptom in migraine 337
intoxication 24
management 49-64
antivertiginous and antiemetic drugs
49-51

central compensation 55-61


drug-modulated compensation

metabolie 24
post-traumatic 342-3
and postural imbalance 25
presyncopal 24
psychosomatic 24
Domperidone 328
Downbeat nystagmus 14,16,27,37,173,
199-205,314

aetiology 202-3,204
clinical syndrome 199-201
nystagmus 200-1
oscillopsia and impaired motion
perception 201
postural imbalance 201
management 204-5
pathomechanism and site of lesions
202

Dreams 17-18
Drop attacks 94,244,388-9
Droperidol SO
Drug-induced vertigo 395-410
cerebellar intoxication 397-400
drugs and eye movements 400-1
ototoxic agents 395-7
Elderly 385-92
cautious senile gait 386-8
dizziness in 389-91
falls in 388-9
physiological ageing of vestibular
system 385-6
age- related changes in eye
movements and vestibulo-ocular
reflexes 385-6
age-related changes in postural sway
and balance 386
Electronystagmography 43
caloric test 44
perilymph fistula 102
rotatory chair and drum 43
vestibular paroxysmia 121
Encephalitis 173
Endolymphatic hydrops 86-7,104
delayed 88, 342
Epidemie vertigo 173,242
Epileptic nystagmus 33,235-7
Episodic vertigo 23- 5
Epley manoeuvre 266,267,268
Ergotamine 328
Erholungsnystagmus 71,90
Ethacrynic acid, ototoxicity 397
Ewald's second law 73
Eye-head synkinesis 107
Eye movements
age-related changes 385-6
drugs and 400-1

58-60

mechanisms of vestibular
compensation 56-8
substitution of vestibular function
60-1

terms and definitions 55-6


surgical treatment 51-2
vestibular exercises and
physiotherapy 52-5
balance training in normal
subjects 52-3
balance training in vestibular
dis orders 53-5

Falls, in elderly 388-9


Familial episodic ataxia 33,173,365-74
aetiology and pathomechanism 370-2
cerebral calcium channelopathy
370-1

effects of acetazolamide 372


potassium channelopathy 370
clinical syndromes 366-9
differential diagnosis 369-70
interictal myokymia 366-7
interictal nystagmus 367-9

499

Index
management 370-1
Familial vestibular areflexia 132
Fatiguability 253
Fencing posture 443
Fentanyl,ototoxicity 398
Finger-pointing test 42
Fixation suppression of vestibulo-ocular
reflex 41
Flicker illumination 426-7
'Floating' labyrinth 99, 104
5-Fluorouracil,ototoxicity 398
Frenzel's glasses 34,68
Frontal disequilibrium 387-8
Frontal gait disorder 388
Functional overlapping 6
Functional specialisation 6
Functional substitution 6
Furosemide,ototoxicity 397
Gait
dis orders 386-8
frontal gait dis order 388
isolated ignition failure 388
psychogenic disorders 461-72
criteria for 462-9
management 469
senile 386-8
Gait ignition failure 388
Galvanic vestibular stimulation 29-31
Gaze-evoked nystagmus 37, 69
Gaze-holding function 6,8,9
Gentamicin
intratympanic therapy 91-2
in Meniere's disease 355-6
ototoxicity 395,399
Gingko biloba 60
Graviceptive pathways 17l-2, 173-4, 179,
182-4
Habituation 56
Haemophilus injluenzae, Meniere's disease
86,377
Halmagyi-Curthoys test 39,129
Haloperidol, ototoxicity 398,399
Head (neck)-extension vertigo 297-8
Head and neck injury 347-50
traumatic otolith vertigo 349
Head-shaking nystagmus 34
Hearing tests
perilymph fistula 102
vestibular paroxysmia 120
Height vertigo 418-23
licences for workers on heights 422
mechanism of 420
physical prevention 422
'visual cliff' phenomenon 422-3
Hennebert sign 101
Hereditary vestibular dis orders 363
in children 378-9
familial periodic ataxia/vertigo
365-74
aetiology and pathomechanism
370-2
clinical syndromes 366-9
management 372-3
Herpes zaster oticus 146-9
Hertwig-Magendie sign 185

Hyperekplexia 244
Hyperventilation syndrome 23,458
Hyperviscosity syndrome 341-2
aetiology and pathomechanism 341
clinical syndrome 341
management 341-2
Hysterical vertigo 458
Iatrogenic vestibular dis orders 355-9
iatrogenic benign paroxysmal
positioning vertigo 357
intratympanic gentamicin in Meniere's
disease 355-6
quinine 356
surgically induced vestibular
dysfunction 356-7
vertebral artery dissection due to
chiropractic neck manipulation,
356
vestibular loss associated with chronic
noise-induced hearing loss
357-8
Ibuprofen 328
Immune-mediated inner ear dis orders
151-5
bilateral vestibulopathy 130
Increased spinal input hypothesis 58
Industrial solvents, ototoxicity 398,399
Internal auditory artery infarction 308-9
Internal representation of verticality 18,
182
Intoxication 24,458
Ion channel defects, 365-73
cerebral calcium channelopathy 370-2
potassium channelopathy 370
Ischaemia
head (neck)-extension 297-8
transient vertebrobasilar ischaemia
296-7
Jugular
diverticle 163
glomus 162
Kanamycin, ototoxicity 389
Klippel-Feil syndrome 145
Labyrinth
computed tomography 143
dis orders 143-70
autoimmune 151-55
Cogan's syndrome 154-5
congenital 143-5
infectious 145-51
acute otitis media 149-51
cholesteatoma 151
herpes zaster oticus 146-9
syphilitic labyrinthitis 151
tuberculous labyrinthitis 151
tumours 155-64
'floating' 99,104-5
magnetic resonance imaging 143-64
Labyrinthine concussion 348
Labyrinthitis 147
syphilitic 151
tuberculous 151

Latent congenital nystagmus 35


Lateropulsion 14,16,243-4
Liberatory manoeuvres 265-8
Light -headedness 23
Linear-rotatory nystagmus 253
Linearvection 409-16
Lisuride 328
Lithium,ototoxicity 398,399
Magnetic resonance imaging
labyrinth and vestibular nerve 143-64
perilymph fistula 102
vestibular cortex 223
vestibular neuritis 70-1
Mal de debarquement syndrome 473
Marijuana,ototoxicity 398,399
Mastoiditis 147
Melanoma 164
Meniere's disease 73,83-98
aetiology and pathomechanism 86-90
aetiology 87-9
delayed endolymphatic hydrops
88
psychosomatic hypothesis 89
vascular hypothesis 88-9
endolymphatic hydrops 86-7
pathophysiology of attacks 89-90
in childhood 377
clinical syndrome 83-5
attacks 83-4
auditory symptoms and signs 84-5
differential diagnosis 85
imaging 85
vestibular function 85
intratympanic gentamicin in 91-2,
355-6
management 90-3
attack -free interval 91
attacks 90-1
intratympanic gentamicin therapy
91-2,355-6
pragmatic therapy 93
surgical treatments 51,92-3
natural course 85-6
surgical treatment 51,92-3
vestibular drop attacks 14,15,26,27,
94
Meningeoma 161
Mercury, ototoxici ty 399
Mesencephalic artery infarction 312-14
Metabolie dis orders of vestibular system
245
Methadone,ototoxicity 398,399
Methysergide 328
Metoclopramide 328
Metoprolol 328
Middle cerebral artery infarction 317-22
Migraine 325-40
aetiology and pathomechanism 326-7
association with other vertigo dis orders
337-8
basilar migraine 327,329-38
benign paroxysmal torticollis in
infancy 336
benign paroxysmal vertigo in
childhood 335-6
benign recurrent vertigo 337
in childhood 370
clinical syndrome 329-33
pathomechanisms 333-5

500

Index

Migraine - cant.
clinical syndrome 326
dizziness and vertigo as facultative
symptoms 337
management 327-8
prophylaxis 328
Minamata disease 399
Mismatch 4-5,483
Mixed canal-otolith vertigo 28-9
Mondini dysplasia 131,147
Motion sickness 485-96
in children 376
clinical syndrome 485
incidence and susceptibility 490
labyrinth function 487
management 491-2
me die al prevention 491-2
physical prevention 491
visual prevention in vehicles 491
nausea and vomiting 485-7
optokinetic 416,417-18
pathomechanisms of 333-5
sensory conflict theory 487-90
space sickness 492-3
symptoms in basilar migraine 334
Multiple sclerosis, central vestibular
syndromes in 244

Mycabacterium tuberculosis 131


Myokymia, interictal 366-7
Naproxen 328
Nausea 4,485-7
see also Motion sickness
Neck reflexes 443

Neisseria meningitidis

131

Neomycin,ototoxicity 389
Neural integrator function 8,9
Noise-induced hearing loss, vestibular loss
associated with 357-8
Non-epileptic cortical vertigo 173
Non-vestibular vertigo syndromes 413
Nucleus ofCajal 176,179,184
Nystagmus 4
arthrokinetic 446-2
central positional 292
cervical vertigo 444
congenital 435-6
convergence- retraction 37
downbeat see Downbeat nystagmus
epileptic 33,235-7
gaze-evoked 37, 69
head-shaking 34
latent congenital 35
linear-rotatory 253
optokinetic 69
with oscillopsia 428-9
positional alcohol vertigo/nystagmus
33,286-7

positional downbeat 291-2


positional glycerol 287-8
positional 'heavy water' 287
positional with macroglobulinaemia
288

positioning 253
post-rotatory 395
rebound 37
spontaneous 34

torsional 173, 193-4,313


upbeat see Upbeat nystagmus
Ocular flutter 36
Ocular tilt reaction

14,16,27,35,177,

179-86

otolith Tullio phenomenon 108


vestibular dis orders
roll plane 173, 179-86
aetiology 184-5
mechanisms of 180-1
natural course and management
185

perceived tilt 181


types of 181-4
Ocular torsion 46, 69
Ocular vertigo 429-30
Off-vertical axis rotation 6
Opsoclonus 36
Optokinetic graviceptive mismatch

102-5

414-16

Optokinetic motion sickness 416,417-18


Optokinetic seizures 237
Orthostatic hypotension 23, 25
OscilJopsia 26-7,430-6
acquired ocular oscillations 432-3
bilateral vestibulopathy 128
deficient vestibulo-ocular reflex 431-2
downbeat nystagmus 200-1
impaired balance 428-9
impaired motion perception 200-1
motion perception in congenital
nystagmus 435-6
normal inhibitory interactions between
self-motion and object-motion
perception 435
pathological binocular impairment of
motion perception 435
physiological impairment of motion
perception 433-4
upbeat nystagmus 206
Otitis externa 147
Otolithic vertigo 29,32,256,347
Otolith Tullio phenomenon 14, 15,33,
106-12,181

clinical history 107


clinical types 107
experimental history 106-7
management 111-12
vestibulospinal reflexes 108-11
Otosclerosis 132
Otosyphilis 147
Ototoxic agents 395-7
Oval window fistula 353-4

experimental perilymph/endolymph
fistulas and endolymphatic
hydrops 104
in childhood 377
clinical syndromes 99-102
differential diagnosis 102
identification 10 1- 2
electronystagmography 102
exploratory tympanotomy 102
hearing tests 102
imaging techniques 102
pressure fistula tests 101
vascular fistula tests 10 1- 2
otolith type 100-1
semicircular canal type 100
fistula of anterior semicircular canal
112

management 105-6
conservative 105-6
surgical 106
post-traumatic 348
Tullio phenomenon 106-12
clinical history 107
clinical types 107
experimental history 106-7
management 111-12
otolith Tullio phenomenon 107-8
vestibulospinal reflexes in 108-11
Peripheral vestibular lesions 55
Peripheral vestibular paroxysmia see
Vestibular paroxysmia
Petrositis 147
Phenothiazines 50
ototoxicity 393
Phenytoin,ototoxicity 397,398,399
Phobie postural vertigo 124,469-78
aetiology and hypothetical mechanism
470-4

Pacinian corpuscle 56
Panic dis order 18,458-9
criteria for panic attack 459
phobie postural vertigo 472-4
Paracetamol 328
Paramedian pontine artery infarction
312-14

Paramedian thalamic infarctions

Paroxysmal dysarthria 33,242


Paroxysmal positional vomiting 293-6
Paroxysmal room-tilt illusion 224-5
Paroxysmal vertigo 29-33,173,242-4
evoked by lateral gaze 242-3
Pendular pseudonystagmus 128
Perceived tilt 181
Perceived vertical 45,189-92
central vestibular versus peripheral
ocular motor lesions 192
historical reports 189
thalamic and cortical astasia 192
versus room tilt 190-2
as vestibular sign 188-90
vestibular paroxysmia 121
Perilymph fistula 27,34,99-115
aetiology and pathomechanisms

315,

316,317

Paraneoplastic syndrome 398


Parieto-insular vestibular cortex 220-1
Paroxysmal ataxia 242

clinical syndrome 469-70


course and treatment 475-8
differential diagnosis 474
Physiological vertigo 483
motion sickness see Motion sickness
Physiotherapy
dizziness 52-5
vestibular neuritis 77-8
Pitch plane 6
signs 169,170
tone imbalance in 171
vestibular dis orders in 205-19
vestibular syndromes in 314
Pizotifen 328

501

Index
Plasticity of vestibular system 55-61
drug-modulated compensation 58-9
substitution of vestibular function
60-1

terms and definitions 55-6


vestibular compensation 56-8
Politzer balloon 34
Position al alcohol vertigo/nystagmus

33,

286-7

Position al downbeat nystagmus 291-2


Positional glycerol nystagmus 287-8
Position al 'heavy water' nystagmus 287
Positional nystagmus 253-4
with macroglobulinaemia 288
Positionalipositioning vertigo 26,253
buoyancy hypo thesis 285-9
Positioning vomiting 293-6
Post-concussion syndrome 103,458
Posterior inferior cerebellar artery
infarction 303,309-12
Posterolateral thalamic infarctions 315,
319

Post-rotatory nystagmus 395,401


Post-traumatic syndrome 348,458
Postural imbalance 27
arthrokinetic nystagmus 446-7
downbeat nystagmus 201
sensory polyneuropathy 447-8
upbeat nystagmus 207
Postural sway
age-related changes 386
nystagmus 428-9
see also Body sway
Posture
and vertigo 68-9,254-6
and vision 423-8
flicker illumination 426-7
moving visual seen es 424
near vision and eye-object distance
425-6

visual acuity 424-5


visual control of fore-aft versus
lateral body sway 426
visual field 427-8
visual stabilization in dark 426
Posturography 44
vestibular paroxysmia 121
Pressure fistula tests 10 1
Prochlorperazine 50
Promethazine 50,491
Propranolol 328
Pseudoataxia 463
Pseudo-Coriolis effect 416-17
Pseudo-Purkinje effect 416-17
Psychiatrie dis orders 455-74
acrophobia 460-1
psychotherapy 461
agoraphobia 459-60
criteria for 459-60
psychotherapy 461
diagnosis of psychogenic vertigo
457-8

organic versus psychiatrie morbidity


456

panic dis order 458-9


criteria for panic attack 459
psychogenic disorders of stance and
gait 461-72
criteria for 462-72

management 469
vestibular dysfunction 456-7
Psychogenic tremors 462-3
Purkinje effects 417
Quinine
ototoxicity 397
side effects 356
Ramsay Hunt syndrome 146-9
Rebound nystagmus 37
Recovery 55
Reflex epilepsy 237
Roll plane 6
signs 169,170
tone imbalance in 170-1
vestibular disorders in 175-203
vestibular syndromes in 307
Rollvection-tilt 414-16
Romberg test 41,462
Room-tilt illusion 33,173,190-2
paroxysmal 224-5
Rotatory seizure 14,173,234
Round window fistula 353-4
Saccades 38,60
Sagittal plane see Pitch plane
Scheibe syndrome 143
Schlagfeld 412, 414, 443
Schwannoma 158-9
Scopolamine 50,76,491
ototoxicity 398
Self-motion perception 225-31
Semicircular canal vertigo 28-9
Semontmanoeuvre 265,266,268
Senile gait 386-8
Sensory polyneuropathy 441, 447
Skew torsion 69,107,171,173,185-9
aiternating skew deviation 187-8
natural course 188-9
types of 186-7
Ski sickness 489
Sleep 17-18
Smooth pursuit 36-7
Somatosensory vertigo 441-51
arthrokinetic nystagmus and selfmotion sensation 446-7
cervical vertigo 441-6
ataxia and nystagmus in
experimental cervical vertigo
444

clinical evidence 444-6


neck afferents and neck reflexes
442-4

central pathways 443


cervico-ocular reflex 443
neck reflexes 443
spatialorientation 442-3
postural imbalance with sensory
polyneuropathy 447-8
Space constancy mechanisms 3,5,6,56
Space phobia 473
Space sickness 492-3
Spatial hemineglect 173,224
Spatialorientation 135,442-3
Spontaneous nystagmus 34

Square-wave jerks 34,36


Stance, psychogenic disorders of 461-8
criteria 462-6
management 463
Stapedius reflex 109
'Startle disease' 234
Static balance 41

Streptococcus pneumoniae 131


Streptococcus suis meningitis 131
Streptomycin,ototoxicity 395,399
Stroke 307-24
anterior inferior cerebellar and internal
auditoryarteries 308-9
basilar, paramedian pontine and
mesencephalic arte ries 312-14
vestibular syndromes in pitch plane
314

vestibular syndromes in roll plane


312-14

cortical infarctions 315-22


peripheral and central vestibular
dis orders 307
thalamic infarctions 314-15
vertebral and posterior inferior
cerebellar arte ries 309-12
Wallenberg's syndrome 309-12
Subjeetive visual vertieal see Perceived
vertieal
Sumatriptan 328
Sustained (rotatory) vertigo 24-5
Sympathetie organ dysfunetion 152
Syncope, classification 25
Syphilitic labyrinthitis 151
Temporal bone, fractures of 164,347-8
Thalamic astasia 14,16,27,173,192,243,
319

Thalamic infarctions

314-15,316,317,

319

Thallium,ototoxicity 398,399
Tinnitus 27
Tobramycin,ototoxieity 389
Toluene,ototoxicity 398
Torsional nystagmus 173, 193-4,313
Tortieollis, benign paroxysmal, in infancy
336

Transient room-tilt illusions 173


Transient vertebrobasilar ischaemia
296-7

Traumatic otolith vertigo 347,349


Traumatic vertigo 345
barotrauma 351-4
altern ob arie vertigo 351-2
blast injury 352
decompression sickness 352
head and neck injury 347-50
traumatic otolith vertigo 349
iatrogenie vestibular disorders 355-9
iatrogenic benign paroxysmal
positioning vertigo 357
intratympanie gentamicin in
Meniere's disease 355-6
quinine 356
surgieally induced vestibular
dysfunction 356-7
vertebral artery disseetion 356
vestibular loss with ehronie noiseindueed hearing loss 357-8

502

Index

Tricyclic antidepressants, ototoxicity


398,399

Tuberculous labyrinthitis 151


Tullio phenomenon see Otolith Tullio
phenomenon
Tumarkin's otolithic crisis 14,15,29,94
Tumours 155-64
acoustic neurinoma 159,160
brain, and vestibular syndromes 244
diverticle of jugular vein 163
jugular glomus 162
melanoma 164
meningeoma 161
schwannoma 158-9
Tympanotomy, exploratory 102
Unsteadiness
bilateral vestibulopathy 127-8
see also Gait
Upbeat nystagmus 173,205-11,314
aetiology 204, 207
clinical syndrome 206-7
nystagmus 205-6
oscillopsia, motion perception and
spatial orientation 206
postural imbalance 207
management 209-10
pathomechanism and site oflesions
207-8

Usher's syndrome

131,143,379

Valproic acid 328


Valsalva manoeuvre 109
Vancomycin,ototoxicity 389
Vascular fistula tests 101-2
Vascular vertigo 303-5
stroke see Stroke
Vertebral artery disseetion 348
due to chiropractic neck manipulation
356

Vertebral artery infarction 309-12


Vertebrobasilar insufficiency 303
Vestibular ageing 385-6
Vestibular ataxia 4
Vestibular atelectasis 27,132
Vestibular autonomie regulation 16-19
neuroanatomie substrates 18-19
Vestibular compensation 56-60
Vestibular cortex 219-31
multimodal sensorimotor
functionl dysfunction 221- 33
paroxysmal room -tilt illusion
224-5

self-motion perception 225-9


spatial hemineglect 224
multiple areas 219-21
no primary vestibular cortex
219-20

parieto-insular vestibular cortex


220-1

Vestibular dis orders 173


brain tumours 244
central brainsteml cerebellar lesions
241-2

central vestibular falls 15-16


clinical classification 169-72
metabolie disorders 245- 6

multiple sclerosis 245


paroxysmal vertigo 29,242-4
pitch plane 199-219
downbeat nystagmus 199-205
aetiology and pathomechanism
202-3

clinical syndrome 199-201


management 203-5
upbeat nystagmus 205-11
aetiology and pathomechanism
207-9

clinical syndrome 206-7


management 209-10
roll plane 175-97
clinical syndrome, 176-8
ocular tilt reaction 179-85
aetiology 184-5
mechanism 180-1
natural course and management
185

perceived tilt 181


types of 181-4
perceived vertical 189-92
central vestibular versus peripheral
ocular motor lesions 192
historical reports 189
thalamic and cortical astasia 192
versus room tilt 190 - 2
as vestibular sign 189-90
skew torsion 185-9
alternating 187-8
diagnostic value 186
natural course 188-9
types of 186-7
topographie diagnostic rules 178-9
torsion al nystagmus 193-4
yaw plane 215-18
combined VOR dysfunction 217
horizontal nystagmus 215-16
Vestibular downbeat syndrome see
Downbeat nystagmus
Vestibular dreaming 18
Vestibular drop attacks 94,388
Vestibular epilepsy 14,15-16,33,171,
173,233-39

epileptic nystagmus 235-7


vestibular seizure 234
differential diagnosis 234-5
management 235
rotatory 234
vestibular versus visual (optokinetic)
seizures 237
'vestibulogenic epilepsy' 237-8
Vestibular exercises 52-5
Vestibular falls 13-16
central 15-16
downbeat nystagmus syndrome 16
lateropulsion in Wallenberg's
syndrome 16
ocular tilt reaction 16
thalamic astasia 16
vestibular epilepsy 15-16
without vertigo 244-5
peripheral 14-15
benign paroxysmal positioning
vertigo 15
bilateral vestibulopathy 15
Meniere's drop attacks 15
otolith Tullio phenomenon 15

vestibular neuritis 14-15


Vestibular migraine 332-3
Vestibular nerve dis orders 143-66
autoimmune 151-5
Cogan's syndrome 154-5
congenital causes 143-5
imaging 143
infeetious eauses 145-51
aeute otitis media 149-51
eholesteatoma 151
herpes zoster otieus 146-9
tumours 155-64
Vestibular neuritis 14-15,67-81,146
aetiology and pathomeehanism 73-6
partial unilateral vestibular loss
73-5

pathomeehanism 73
site oflesion 76
viral!vaseular aetiology 75-6
eentral brainstem lesions mimieking
241-2

in ehildhood 376
clinieal syndrome 67-73
ealoric testing 69-70
differential diagnosis 72-3
eye movements 69
high-frequeney defect ofVOR 71-2
MR imaging 70-1
spontaneous course 71
vertigo and posture 68-9
management 76-8
Vestibular paroxysmia 14,33,117-26
aetiology and pathomeehanisms 122
alternating vestibular nerve paroxysmia
and failure 124-5
clinieal syndrome 117 - 22
auditory symptoms and tests 118
ease reports 118-21
differential diagnosis 121-2
eleetronystagmography 119
posturography 119
subj eetive visual vertieal 119
management 122-4
Vestibular plastieity 55-61
drug-modulated compensation 58-9
substitution of vestibular funetion
60-1

terms and definitions 55-6


vestibular eompensation 56-8
Vestibular seizure 234
differential diagnosis 234-5
management 235
rotatory 234
Vestibular substitution 60-1
Vestibular suppressants 49
Vestibular upbeat syndrome see Upbeat
nystagmus
Vestibular vertigo syndromes 3-9
mismateh coneept 4-7
signs and symptoms 4
vestibulo-collie reflex 10
VOR-mediation of pereeption and
postural adjustments 9-10
Vestibulo-collie reflex 10
Vestibulogenie epilepsy 237-8
Vestibulo-oeular adaptation 60
Vestibulo-oeular reflex 5-10,39-41,169
age-related ehanges 385-6
bedside test 39,129

503

Index
bilateral vestibulopathy 127
defective 431-2
neuronal network 7-9
perception and postural adjustments
9-10

transmitters of 58-9
visual fixation suppression 41
Vestibulospinal reflexes 10-12
otolith Tullio phenomenon 108-ll
Visual acuity 424-5
Visual cliff phenomenon 422-3
Visual field 427-8
Visual (optokinetic) seizures 243
Visual vertigo 409-39,473
circularvection and linearvection
409-16

psychophysics of circularvection
411-l3

rollvection-tilt 414-16
visual-vestibular interaction 4l3-14
eye movements, oculomotor disorders
and postural balance 428-30
extraocular muscle paresis 429-30
nystagmus with oscillopsia 428-9
optokinetic motion sickness 417-18
oscillopsia 430-6

acquired ocular oscillations 432-3


deficient vestibulo-ocular reflex
431-2

inhibitory interactions between selfmotion and object-motion


perception 435
motion perception in congenital
nystagmus 435-6
pathological binocular impairment of
motion perception 435
physiological impairment of motion
perception 433-4
physiological height vertigo and
postural balance 418-23
licence for workers on heights 422
physical prevention 422
'visual cliff' phenomenon 422-3
vision and posture 423-8
flicker illumination 426-7
moving visual scenes 424
ne ar vision and eye-object distance
425-6

visual acuity 424-5


visual control of fore-aft versus
lateral body sway 426
visual field 427-8

visual stabilization in dark 426


visual pseudo-Coriolis effect and
pseudo-Purkinje effect 416-17
Visual-vestibular interaction 225-9,
4l3-14

Visual-vestibular mismatch 487-90


Volvular epilepsy 16,173,234
Vomiting 485-7
paroxysmal positioning 293-6
see also Motion sickness
Waardenburg's syndrome l31
Waldenstrm's disease 288
'Walking on ice' 462-4
Wallenberg's syndrome 6,52,174,176,
184,243-4,309-12
14, 16, 244

lateropulsion in

Yaw plane 6
signs 169,170
tone imbalance in 170,314
vestibular dis orders in 215-18

También podría gustarte