Está en la página 1de 24

Archaeometry 50, 2 (2008) 313336

doi: 10.1111/j.1475-4754.2008.00388.x

DETECTING TRENDS IN THE PREDICTION OF THE


BURIED PAST: A REVIEW OF GEOPHYSICAL
TECHNIQUES IN ARCHAEOLOGY*

Oxford,

0003-813X
Archaeometry
ARCH
Original
XXX
A
C.
*Received
review
University
Gaffney
UK
Articles
of
11of
geophysical
October
Oxford,
2007;
2008techniques
accepted 7 December
in archaeology
2007
Blackwell
Publishing
Ltd

C. GAFFNEY
Archaeological Sciences,
Division of Archaeological, Geographical and Environmental Sciences,
University of Bradford, Bradford, West Yorkshire BD7 1DP, UK

Geophysical survey techniques are a highly visible part of the scientific toolkit that is now
used by archaeologists. In this paper, the history of the use of geophysical techniques in
archaeology will be discussed, as will significant research themes associated with the most
widely used prospecting devices. It is apparent that while the use of geophysical techniques
is at an all-time high, there are many key areas where prospecting is rapidly developing.
Some of the advances relate to fundamental aspects of the techniques, while others dictate
how we undertake survey in the future. There is a movement away from pre-gridded survey
areas towards real-time GPS for navigation. This allows greater integration, or fusion, of
disparate data sources using visualization techniques derived from associated disciplines.
The analysis of landscapes has become a major component of the application of new technology
and there are many challenges to be tackled, including how to analyse and interpret
significant archaeology within large-scale, data-rich, multi-technique investigations. The
reflective nature of the review acknowledges the important role of Archaeometry in the
development of archaeological geophysics.
KEYWORDS: MAGNETOMETRY, EARTH RESISTANCE, RESISTIVITY, GPR,
ELECTROMAGNETIC, VISUALIZATION, GPS, LANDSCAPE
8 University of Oxford, 2008

INTRODUCTION

The use of geophysical techniques to identify buried archaeology can be charted a number of
years before the publication of Archaeometry. While it was commonly held that the birth of
archaeological geophysics was at Dorchester on Thames in the period immediately after the
Second World War (Atkinson 1953), recent historical probing has revealed significant,
deeper strata that confirm that an electrical resistance survey undertaken in the USA predates
that time (Bevan 2001).
However, the main thrust of the early publications in Archaeometry was linked to magnetic
prospecting; typically for the period, the archaeological framework was only loosely wrapped
around the scientific endeavour. The short account of a survey using a new proton magnetometer
at Water Newton, published by Martin Aitken (1958) in the first volume of Archaeometry,
compressed the huge effort that had actually gone in to the designing, building and testing of
the kit and the subsequent success of locating a previously unknown kiln. A wry commentary
on this formative survey was later published by Aitken (1986), in which he suggested that the
real archaeological significance of the results was not immediately evident. Aitken admitted to
a feeling of disappointment when rubbish-filled pits were detected rather than pottery kilns.
*Received 11 October 2007; accepted 7 December 2007
University of Oxford, 2008

314

C. Gaffney

In the second issue of Archaeometry, a note was published by Fowler (1959) stating the archaeological case for the identification of rubbish pits, and a whole new rationale was developed,
away from anomalies of thermoremanent origin towards the more subtle magnetically induced
anomalies that are considerably more common in the archaeological record. It could be argued
that a lack of archaeological context was to hinder the advance of geophysical techniques for
many years to come, but it should not be assumed that the barriers were simply a result of
misunderstanding by the scientists who were at the forefront of the development of new
techniques. In 1956, the first edition of Hoskins influential book on landscapes was published; in
the introduction of the second edition (1976), he wrote: The student . . . faces at times the
possibility of underground evidence . . . the visible landscape offers us enough stimulus and
pleasure without the uncertainty of what might lie underneath.
In some ways, the end of the 1970s was the nadir for geophysical prospecting for archaeological purposes. There was a feeling that many of the big problems had been solved and that
greater fundamental problems existed elsewhere in archaeological science. Although many
articles on prospecting graced the pages of Archaeometry in the 1960s, by the 1970s the flow
of material had reduced to a trickle. In part, that can be attributed to the changing interests of
the burgeoning research group at Oxford but, more importantly, it reflects the impact during
the late 1960s of the specialist journal Prospezioni Archeologiche (PA), published by the
Lerici Foundation in Italy. That journal was heavily influenced by Richard Linington, who
previously had been based at the research laboratory at Oxford. Following from Archaeometry,
PA investigated the broad spectrum that now encapsulates geophysical surveying, including
modelling, display and all aspects of theory and practice. A brief perusal of the index of that
journal will identify all the major players from the embryonic period of the discipline
Aitken, Aspinall, Clark, Hesse, Linnington, Scollar and Tite are amongst them. The journal
published articles on essentially the whole host of non-destructive techniques that were
available. However, as publication of PA became erratic, so an apparent slump in the fortunes
of geophysical surveying occurred (see Gaffney and Gater 2003): Linford (2006) considers
this to be more of a period of retrenchment towards applications rather than fundamental
research. By the late 1980s this was indeed true, as instrument and software design became
increasingly robust and the opportunities for prospecting rather than contextualizing became
more apparent.
Moving closer to the present, the day-to-day use of geophysical techniques, and more
pertinently remote sensing or prospecting as a whole, is remarkably buoyant. This begs the
following questions: What were the pivotal points in this resurgence, and what makes
geophysical techniques more appropriate to archaeology as it is practised now as opposed to the
1970s? Essentially, archaeology has a need for fast and accurate mapping of buried remains,
primarily for planning purposes, but it also requires an appropriate alternative to costly and
destructive invasive techniques. Not only has excavation become increasingly expensive; in
some parts of the world, this traditional archaeological tool requires permits that are difficult
to obtain. Perhaps more surprisingly for some readers, there is a reluctance to dig that is based
on conservation and ethical grounds. Ultimately, geophysical techniques have partly filled this
gap, due to the increasing reliability of the instruments and the explosion in computing power.
Kelly et al. (1984) published in Archaeometry the first direct coupling for archaeological
purposes of a resistance meter to a portable computer. Physical links to expensive field
computers were to be used for only a short time, as the next generation of equipment contained
dedicated loggers or on-board memory that allowed downloading for processing at a later
time. Ironically, detection devices are again now occasionally linked to notebooks as part of
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

A review of geophysical techniques in archaeology

315

Figure 1 The rise in use of geophysical techniques in the UK: it is likely that the increase in the past decade
is mirrored elsewhere in the world. The increase can be linked to the proliferation of commercial groups working
on development-led projects. In the UK, magnetometry is the most used technique and while that is probably
true globally, there are some countries where earth resistance or GPR rival magnetic investigation
(after Gaffney and Gater 2003).

GPS navigation systems. In the UK, the rise in the use of geophysical techniques has been
charted as shown in Figure 1. Although there are local factors that revolve around the use of
non-invasive techniques for planning purposes, there is reason to believe that such devices are
commonly, and increasingly, used throughout the world.
During the 1990s, an unprecedented amount of survey work was undertaken: it is true that
the majority was to enhance the archaeological record prior to a decision to approve or reject
a proposed development (infrastructure, housing, mineral extraction etc.). However, implicit in
this use was the belief that geophysical techniques could do more than provide a context for
an excavation. Prior to this point in time, the accusation could be levelled that geophysics was
simply glorified wall-following. That accusation appeared in a volume that linked geophysics
to landscapes for the first time (Spoerry 1992, 2): the agenda that was driven by some of the
contributors was unachievable in the short term, but is now embedded in the rationale for
terrestrial-derived data. This is particularly true for techniques that have been used to map
areas (e.g., resistance and magnetometry) rather than those that have traditionally been used to
investigate vertical sections of the Earth (e.g., electrical imaging or seismic investigation). By
the end of the 1990s, the methodology for handheld devices had become extremely efficient,
with 12 ha covered in a single day.
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

316

C. Gaffney
Table 1 The most frequently used geophysical techniques for terrestrial investigation: active indicates
that the technique induces the phenomenon that is measured, while passive techniques measure what is
there naturally (an update from Gaffney and Gater 2003)

Method

Active or passive

Frequency of use

Magnetometry
Electrical resistance/resistivity
Ground penetrating radar
Electromagnetic
Magnetic susceptibility
Metal detectors
Seismic
Microgravity
Induced polarization
Self-potential
Thermal

Passive
Active
Active
Active
Active
Active
Active
Passive
Active
Passive
Passive

High
High
Highmiddle
Middle
Middle
Low
Low
Low
Low
Low
Low

GEOPHYSICAL TECHNIQUES FOR LOCATING AND DELIMITING THE


ARCHAEOLOGICAL RESOURCE

Although the majority of the surveys for archaeological purposes use magnetic properties to
detect buried archaeology, there are a number of techniques that are used quite frequently for
this purpose (see Table 1). The basis for each technique is beyond the scope of this review, but
there are a number of general works that can be consulted, which cover both the theoretical
and practical scope of the subject (e.g., Hesse 1966; Wynn 1986; Clark 1990; Scollar et al.
1990; Neubauer 2001; Linford 2006). There is, however, considerable merit in charting the
present state of some of the methods that have made, in archaeological terms, the greatest
impact.
Magnetometry
As outlined above, magnetometry is the most common form of geophysical investigation tool
used for archaeological purposes. The majority of the instruments that are in use today stem
from either fluxgate (Alldred 1964) or optically pumped technology (Ralph 1964) and, as
such, are well known both in design and archaeological output. The proton systems used in the
earliest surveys quickly fell out of fashion, as they were not as efficient for ground coverage
as either fluxgate or optically pumped systems. However, there are inherent differences
between the two predominant techniques; the former measure the vertical component of the
Earths magnetic field, while the latter respond to the total field. They are rightly
regarded as the workhorses of prospection, due to the rapidity of data collection and the
general applicability of the technique to shallow investigation. Indeed, many of the great
surveys have been collected using magnetometry, from the small-scale site anomaly characterized
by the work by Aitken (1958) at Water Newton, through detailed analysis of sites across the
world (see Fig. 2), complete cities (Gaffney et al. 2000; Keay and Millet 2000) and on towards
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

A review of geophysical techniques in archaeology

317

Figure 2 A fluxgate gradiometer survey from Tell el-Balamun, using a Geoscan Research FM36 fluxgate
gradiometer. Prior to the survey, large parts of the Temple enclosure were blank. Additional buildings and even
a previously unknown temple have been identified in the magnetometer data. This survey illustrates the
level of detail that can be mapped in good-quality data. Minimal processing has been undertaken on this data
(image courtesy of T. Herbich).

landscapes of many varieties (Sutherland and Schmidt 2003; Powlesland et al. 2006). Given
the immense scale of many of the latest magnetometer surveys, it is perhaps surprising that the
majority have been collected using handheld single sensors. However, this situation has
changed recently; dual and even greater multi-sensor systems are now relatively common.
The progression from single to multi-sensor systems, and the provision of a non-magnetic
cart, can be seen in Becker (2001). Multi-sensor, cart-based systems have been a feature of
caesium vapour measurements over the past decade or so, and have provided many scientific
and archaeological highlights (Neubauer 2001). The benefits of using such systems are
immense: data that are largely free from periodic errors, rapid data collection and, assuming
that sensors are mounted close together, very high data density. A summary of the theoretical
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

318

C. Gaffney

and practical strengths and weaknesses of magnetometers can be found in Aspinall et al.
(in press), while additional information can be found on the sensitivity of the most recent fluxgate
and caesium vapour sensors in Linford et al. (2007). It is clear that the archaeological
geophysicist now has real choices with these instruments: vertical or total field, single sensor,
gradiometer or differential mode, handheld or non-magnetic cart, human- or vehicle-powered.
Crucially, the question is not which sensor is best, but which is most appropriate. These
significant advances are invaluable if magnetometry is to be used in all parts of the globe,
especially away from the mid-latitude zone (Tite 1966). Additionally, what is clear is that there
is a necessity and an ability to measure considerably smaller responses than Aitken and
colleagues ever thought possible (for the first picotesla measurements, see Becker 1995). What
is of interest, in the context of this review, is that the drive behind the need for this increase in
sensitivity relates to archaeological problems rather than a change in technology per se. In
particular, there is considerable concern over the limits of detection for particular magnetometer
sensors/configurations, especially in areas of potential low contrast that are often found, for
example, on ritual sites or over deeply buried targets.
Given this scenario, one exciting development is the use of SQUID (Superconducting Quantum
Interference Device) based systems for highly sensitive magnetic field measurements (Chwala
et al. 2001). Originally, these devices were used for laboratory measurement of weak magnetism
(Walton 1977) and significant technological barriers have been overcome to produce an
operational instrument that can be used for prospecting purposes. As the base length of the
integrated SQUID is only 4 cm, the sensors measure true gradients, and values in the order of
a few fT (1015 nT) should be achievable. The main limitations of SQUIDs are that they must
be kept very cold for superconductivity to occur, and that the short base line is essential to
suppress both the Earths field and that associated with any mechanical device that is used to
pull the sensors. Conventionally, so-called Low Temperature (L.T.S.) measurements are collected
(~ 4 K, liquid helium cryostat), although replacing the sensor material with complex compounds allows superconductivity to occur at liquid nitrogen temperatures (High Temperature,
or H.T.S.; Chwala et al. 2003). A viable L.T.S. SQUID field system is now in use (Schultze
et al. 2007). In addition to the increased sensitivity, sampling rates are huge, allowing the sensors
to be pulled at high speeds (greater than 30 km h1) behind a car (Fig. 3): great care must be
taken to correct for the relative position of the vehicle, as well as the pitch and roll of the
sensor array (Schultze et al. in press). Although it true that this system easily outperforms the
most sensitive total field sensors that are pulled on a cart, there are still some fundamental
issues to be resolved when considering archaeological applications. Logistically, the use of a
cryostat and the resulting ongoing costs may be problematic, but of greater importance is the
configuration of the sensors. The present sensors are set to measure the vertical gradient of the
horizontal component of the Earths field (dBhorizontal/dz), and a result of measuring this true
gradient is that the response over a buried ditch looks like a derivative of the total field
response (Schultze et al. in press). Archaeologically, this leads to more challenging data
sets, and research will be required to model the potential outputs if the hoped-for advance in
interpretation is to be achieved.
GPS for navigation and location is a common theme in many prospection applications.
It should be noted that it is not used solely with rapid, vehicle-powered systems, but it is
also incorporated in some human-propelled sensor arrays (Fig. 4). The benefits potentially
include a grid-free strategy and precise data location that increases the data quality and
reduces processing. However, potential damage due to the use of heavy vehicles should not be
overlooked when considering prospecting on culturally or environmentally sensitive sites.
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

A review of geophysical techniques in archaeology

319

Figure 3 A SQUID sensor array pulled behind a car. As a result of high sampling rates and real-time GPS
navigation, survey speeds in excess of 30 km h1 can be achieved.

Figure 4 The photograph on the left shows a Foerster Ferrex three-fluxgate sensor array on a non-magnetic cart.
No grid is set out before survey, as the navigation and the actual location of measurements are controlled by
real-time GPS. On the right is a magnetometer greyscale image draped over the topographic data. Both data
sets were collected simultaneously within each magnetometer survey area. The data was collected on behalf of
the Cyrenaica Archaeological Project, with assistance from V. Gaffney.

University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

320

C. Gaffney

Earth resistance/resistivity survey


Traditionally, resistivity surveying in geophysics has been undertaken either to map areas or
investigate vertical sections through the ground. For archaeological purposes, area surveying
has been the dominant method, with the production and analysis of pseudo-sections usually
regarded as a problem for deep geophysics. As with a number of techniques, notably groundpenetrating radar (GPR) and seismic survey, this division has been blurred in recent years, as
it is now possible to create horizontal maps from closely packed vertical sections.
Although the early issues of Archaeometry and the journal Archaeological Prospection are
littered with (literally) ground-breaking studies using novel resistance arrays, they were largely
based around those used in traditional geological situations; that is, Wenner and Double Dipole.
Although not the first to suggest the Twin-Probe, Aspinall became associated with this array
(Aspinall and Lynam 1970) and for many reasons this became the first-choice resistance array
for archaeological prospectingspeed and simplification of anomaly response being among
the most important aspects. Although the Twin has provided many examples of near-perfect
responses from archaeological sites (see Fig. 5), the archaeological world became over-reliant
on this array. However, recent research has re-evaluated some of the previously used arrays
and has re-invigorated area survey. While Twin-Probe, or a derivative of it, is still the most
commonly used array, it is likely to be used because it is appropriate rather than because it is
the only option available. It is not without irony that Aspinall, who championed the TwinProbe, has been at the forefront of array re-evaluation. While the use of the Schlumberger
(Aspinall and Gaffney 2001) and the square array (Aspinall and Saunders 2005) have become
more common, these and other arrays have been included in multi-technique investigations.
Significant increases in on-board computing power have led to an advance in area survey.
Walker (2000) reported the use of a multiplexer, based on the popular Geoscan Research
RM15, that allowed many different resistance readings to be collected at one point. Thus, area
maps of resistance could be drawn for different depths, allowing depth extent to be analysed
for specific features (Fig. 6). Although Walker largely used the Twin-Probe in his analysis,
other arrays can be programmed. Wide-spaced remote probes are often used, commonly
described as the Pole-Pole (Dabas et al. 2000), thereby reducing the ambiguity when converting
to resistivity, which is essential when modelling and inversion is undertaken (Papadopoulos
et al. 2006).
A continuous research thread for archaeological purposes has been an attempt to collect
earth resistance data automatically, using wheels, rotating discs and liquids as electrodes. The
RATEAU system has a long history (Dabas et al. 1994), and its most recent incarnation is a
major contribution to shallow investigation (Dabas et al. 2005; Dabas in press). As can be seen
in Figure 7, the system is GPS controlled, thereby breaking away from the shackles of a small
grid. It is pulled by a quad bike and can cover many hectares per day, making ground coverage
similar to that of a magnetic survey. Additionally, three different depths can be measured,
which means that features such as large ditches, which at certain times of the year can be
invisible (Hesse 1966; Clark 1990), may be recognized, as the system can map moisture
contrasts that exist at different depths. The present version was not made for archaeological
surveying, but for agricultural soil investigation. This illustrates the interlacing of many
near-surface zones that now require analysis, and this will be a fertile ground for future crossdiscipline collaboration. A final point here is that this system is not suitable for Twin-Probe
survey, and thereby demonstrates the move away from this array. However, the images that are
possible with the so-called ARP array (Dabas in press) are highly compatible with other less rapid
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

A review of geophysical techniques in archaeology

321

Figure 5 The Twin-Probe resistance survey over Tockenham Roman villa. Comparative magnetometer data are
also shown to illustrate the response to different features (from Gaffney and Gater 2003).

arrays (see Fig. 8). Evidently, it is not always possible to get a quad bike on to an archaeological
site; it may be too remote, the area too small or the surface too fragile for a heavy vehicle.
In such cases a square array, grid-based wheeled system has been tested (Walker et al. 2005)
and can be modified to simultaneously collect fluxgate gradiometer data (Walker and Linford
2006). Whichever method is used, it is clear that the scope of area resistivity surveying has
advanced considerably in the past decade and that this technique is likely to play an increasing
role in archaeological/environmental research.
The investigation of resistivity with depth has a long tradition in geological geophysics,
but until recently this form of analysis was rarely undertaken for archaeological purposes.
However, production of pseudosections using long linear arrays of multiplexed electrodes, and
the subsequent analysis using standard approaches in geological geophysics, have become more
common. The analyses are usually in the form of tomographic study or by iterative inversions
of the field data (Noel and Xu 1991; Griffiths and Barker 1994; Loke and Barker 1996). From
the archaeological perspective, the problems that are tackled are often very specific, such as
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

322

C. Gaffney

Figure 6 Data collected using the RM15 Multiplexer, illustrating maps at different depths and resolutions,
depending on the geometry and separation of the probes (image courtesy of Roger Walker).

the investigations of mounds. On occasions, the traverses cover very steep slopes and accurate
topographic corrections are required before anomaly shapes can be identified (Tonkov and
Loke 2006; Astin et al. 2007). It is clear that the methods that are used follow those established in other disciplines, but it is the applications in archaeology that are of interest, because
they stretch the interpretative skills of the geophysicist. For example, Berge and Drahor (2007)
produced a three-dimensional synthetic model to invert data collected over a tell-type settlement
in Turkey, which is a very complex problem to tackle from the surface and required a dense
grid of probes (1 1 m) over the 20 20 m survey area. It is evident that such small-scale
investigations may become common only if the field and computational time required produce
clear benefits to those preparing to excavate a site. Again, there is a need to model smaller
features that archaeologists are keen to locate, so that studies cannot simply be accused of wall
following. As opposed to the examination of small features, a cross-disciplinary approach is
apparent when one considers gross changes in archaeologically significant sediments at the
landscape level (see Fig. 9).
Other problems that are important in the study of historical sites in particular are the use of
electrodes that do not damage the surface that is being investigated. Tsokas et al. (in press)
report on the use of flat-based electrodes that use conducting gel to create a good contact with
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

A review of geophysical techniques in archaeology

323

Figure 7 ARP data collection using real-time GPS to navigate. A notebook in front of the driver displays the
present location of the array and the previously collected data pointsthere are three different depths of prospection
(image courtesy of M. Dabas).

stone floors. This allows a detailed survey in confined areas where, for example, a GPR survey
may be less successful due to above-ground reflectors (Nuzzo 2005). There are also applications
for explicitly non-destructive electrodes on the exterior of large-scale monuments, such as
the Acropolis in Athens. As can be seen in Figure 10, experienced climbers were required to
position the electrodes: wire contacts were embedded into wet betonite and taped in place.
This case study is of additional interest because the surveyors had a number of options regarding
data collection and, hence, analysis (Fig. 11). Effectively, they initially analysed the data
assuming borehole to surface electrical resistivity tomography (ERT), as well as modifying
cross-wall data to allow a 3-D volume to be created (Tsokas et al. 2006). This was a very specific
application, but evidently the process and the novel electrodes are likely to be of use on other
sensitive archaeological structures. Capacitive coupled arrays (Panissod et al. 1998) evidently
are of some interest, but due to problems of coupling on uneven ground it is unlikely that they
can produce results that are coherent at the small scale that archaeologists demand, although
there is evidently scope for deeper archaeological or environmental targets (Linford 2006).
Ground-penetrating radar (GPR)
GPR is one of the few new geophysical techniques that have sprung up since Archaeometry
first appeared in print. Although the military use of radar is well known, it was not until the
mid-1970s that archaeological applications became apparent (Vickers and Dolphin 1975). The
resolving power of GPR, in terms of archaeological targets, became an increasingly important
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

324

C. Gaffney

Figure 8 A resistivity survey of the Fanum (temple area) at Vieil-Evreux, France. Data is collected using the
ARP system and the survey is about 1 ha in area: stone structures are revealed as high resistivity (white)
(image courtesy of M. Dabas).

factor in the acceptance of the technique. Factors such as wavelength and the associated footprint have been well researched for the present generation of pulse radar systems, normally in
the 100 MHz to 1 GHz range (Conyers and Goodman 1997; Leckebusch 2003; Conyers 2004).
Partly as a result, the analysis of GPR data has moved away from single traverses towards
dense grids, often with an inter-traverse distance of 0.5 m or less, and with a maximum in-line
step of 0.05 m. Conventionally, GPR has been regarded as a slow, and therefore expensive,
approach to archaeological investigations. However, modern systems are often carried on
lightweight wheeled carts/sledges, and occasionally two or more antennas are fixed together
and pulled behind a motorized vehicle, which navigates using GPS (Leckebusch 2005). Multichannel GPR systems are also available for large survey areas.
Dense grids of vertical profile data are required to produce time or depth slice maps,
which have become the main delivery of GPR data. However, once in a three-dimensional
cube, the data can be analysed and visualized in many different ways. This includes multifacet slicing, amplitude surface rendering and animated fly-throughs, and it leans heavily on
the experience from related signal processing used in seismic studies (Fig. 12). Some of the
commercial software has brought new processing and visualization options to the archaeological
geophysicist, but there are still many fundamental challenges to ensure that an imaged anomaly
equates with buried archaeology. For example, although migration (collapsing the anomaly
hyperbola to its apex) can reduce an anomaly towards its target shape, there are concerns
about the variation in velocity, both laterally and vertically, through a survey area and through
a target itself. This can lead to gross errors that reduce the efficiency of the technique and
increase levels of doubt among the archaeological community. Leckebusch (2007) has provided
one solution to this problem, but it is likely that other approaches will be forthcoming.
Although the majority of the GPR surveys that are undertaken are to investigate targets
buried in the soil, high-frequency antennas are often used for investigating cavities in walls.
As an alternative, Somers et al. (2005) reported in Archaeometry a trial of a continuous fixed
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

A review of geophysical techniques in archaeology

325

Figure 9 An analysis of a former landscape using a resistivity section and two area conductivity surveys
(using EM devices sensing to different depths). (Image courtesy of M. Batessee Bates et al. (2007)
for further details. Copyright 2007 John Wiley & Sons Limited. Reproduced with permission.)

radio frequency system, where the transmitter is separated (down a borehole or on the far side
of a wall) from a passive recorder that moves over a grid at the surface. There is also interest
in developing stepped frequency and frequency-modulated, continuous-wave GPR systems
(Kamei et al. 2000), as well as multi-fold (or multi-offset) collection strategies. Multi-fold
surveys have indicated increased signal-to-noise, allowing more subtle near-surface anomalies
to be identified and potentially indicating dipping strata in a more convincing manner (Pipan
et al. 1999). Recently, Berard and Maillol (in press) have indicated that multi-offset surveying
can be used to investigate sites of great complexity, indicating the base of archaeological
features more effectively than traditional collection strategies. The speed of collection must be
of concern, but any survey strategy that increases the archaeological information prior to
costly excavation must be explored.
Electromagnetic survey (EM)
For archaeological purposes, the majority of electromagnetic surveys use the so-called slingram
set up with a continuous-wave, low-frequency transmitterreceiver pair (Tabbagh 1986a).
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

326

C. Gaffney

Figure 10 A climber scaling the south wall of the Acropolis hill to fix non-destructive electrodes for electrical
resistivity tomography (ERT). The survey was undertaken to establish the construction techniques and the stability
of the wall for conservation purposes (image courtesy of G. Tsokas).

Nearly all of the reported surveys utilize the Geonics EM38 system (14.2 kHz), either for
conductivity mapping (quadrature phase) or magnetic susceptibility (in-phase), but with the
most advanced version of this instrument both parameters can be measured at the same time
and at two different depths. Despite the long history of use for deep prospecting (see Fig. 9)
and the fact that both magnetic and conductivity parameters can be estimated, this technique
remains underused. Part of the problem is that there is ambiguity between the two phases over
high susceptibility and high conductivity soils. As reported in Archaeometry, there are further
considerations to be made, depending on the orientation of the two coils (Tabbagh 1986b),
and field tests have added to the uncertainty of the archaeological interpretation of this data
(Cole et al. 1995; Linford 1998). Benech and Marmet (1999) illustrate these problems using
both theoretical and model field studies, but conclude that the slingram is more effective at
archaeologically significant depths than single-sensor instruments (see below).
Magnetic susceptibility (MS) can be measured using a number of field instruments. Some,
such as the EM38 described above, measure relatively large volumes of earth, while a singlesensor system such as the Bartington uses smaller coils and analyses smaller volumes. From
the surface, as Benech and Marmet (1999) illustrate, slingram instruments are more effective
at locating buried features, but that is only one aspect of the problems that can be resolved
using MS. In fact, the applications for MS are varied, from micro-surveys based on small
excavation units to landscape analysis. A review of MS in the USA by Dalan (2008) has
recently been published and illustrates the breadth of applications. In particular, as a link
between buried feature and magnetometer anomaly, the accurate measurement of MS in three
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

A review of geophysical techniques in archaeology

327

Figure 11 Two-dimensional ERT sections were collected at the south Acropolis wall using different measuring
modes: wall-to-surface (top left), vertical on the wall (top right) and horizontal on the wall surface and on the surface
of the hill (bottom left), while in some open areas it was possible to collect cross-wall ERT data (bottom right)
(image courtesy of G. Tsokas).

dimensions cannot be underestimated. Although laboratory-based analysis is beyond the scope


of this review, it has had significant effects on the understanding of above-ground surveying.
The presence of magnetic bacteria (Fassbinder et al. 1990; Fassbinder and Stanjek 1993) as a
significant contribution to enhanced magnetic levels associated with rotted wood has brought
increased awareness of post-built structures and timber circles in magnetometer data (Neubauer
2001; David et al. 2004). In the commercial arena, wide-spaced (510 m) MS readings are
often measured over large areas, and are interpreted with the rule of thumb that enhanced levels
of MS can be equated with previous human occupation. These hotspots are then targeted for
detailed surveying with another technique. Certainly, in areas of uniform geology, topsoil and
land use, this link can be used as a prospecting tool for long-lived or intense occupation sites.
However, there are very few published examples where the method is tested (see Aubrey et al.
2001; Gaffney and Gater 2003, 96101), so the general applicability still needs more research,
especially for seasonal sites or those that do not contain habitation or industrial remains.
In recent years, there has been an increase in the number of published EM surveys (e.g.,
Witten et al. 2000; Perssona and Olofsson 2004; Venter et al. 2006; Johnson 2006 and the
examples therein). Given the theoretical complications expressed by the authors quoted in the
above paragraphs, there must be mixed feelings in this apparent upsurge. As a community we
need more data, but we also need to be considerably more critical of environmental context
and coil orientation and separation if we are to advance archaeological interpretation of the
data. Many of the published examples are in multi-component strategies, so analyses similar
to those by Cole et al. (1995), Linford (1998) and Benech and Marmet (1999) should be
possible. In short, we have a technique that potentially offers a great deal in conditions where
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

328

C. Gaffney

Figure 12 A 3-D GPR volume from the villa of the Roman Emperor Traianos, Italy. The image is created from
0.5 m interval profiles collected on a regular grid. Clear evidence can be seen for multi-phase use of the site
(image courtesy of D. Goodman).

some of the other techniques are lacking, but the response from each survey area requires
careful and expert analysis. Multi-frequency EM systems have started to appear on the market,
but their use in the archaeological zone is yet to be fully explored.
VISUALIZATION

The ability to visualize data in a more meaningful manner is one of the evident bonuses of
digital data sets. A brief review of the early academic papers in Archaeometry shows that an
ability to draw graphs and hand contour data was a prerequisite for publication. Few researchers
were able to take advantage of the then emerging computing facilities, although those that did,
such as Richard Linnington (e.g., Linington 1964) and Irwin Scollar (e.g., Scollar 1968)
contributed much. However, since the digital capture of data became more frequent in the late
1980s and, probably more important, given the exponential rise in cheap computing power
over the intervening period, the opportunities to analyse and visualize have increased tremendously. Although, traditionally, geophysical data has been processed and displayed in bespoke
software, the inherent digital nature means that visualization can occur in a variety of platforms,
from general display packages, through GIS to relatively seamless integration into specialist
software. The last option borrows heavily on earth science or medical imaging standards, but
benefits from significant and ongoing research in other subject areas (Gaffney and Gaffney
2006).
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

A review of geophysical techniques in archaeology

329

Perhaps one of the most positive aspects of the visualization of data is the acceptance that
geophysical data do not exist in splendid isolation, but can, and indeed must, connect with
other digital data sets. As such, geophysical data can link with all forms of prospection data
that can be mapped on a common grid, and this facilitates analysis and interpretation of
archaeological problems via GIS (Chapman 2006). While it is true that GIS can be regarded
as a basic tool, the potential from the geophysical point of view is considerable, as the
environment is flexible in terms of integrated and enhanced interpretation. The spatial attributes
of features and sites are increasingly important as archaeologists demand more complex models
of the past (Fig. 13) and GIS is a valuable, if not a fundamental, tool in this regard (Kvamme
2003; Neubauer 2004). As the reliance on geophysical data increases for landscape analysis,
the value of GIS is evident if substantive archaeological questions are to be answered (see
Powlesland et al. 2006).
Early experiments in image combination, after normalization of magnetic and resistivity
data, demonstrated the potential interpretative benefit of the reduction and classification of
geophysical data (Neubauer and Eder-Hinterleitner 1997) and additional attempts at mathematical integration have continued since then (e.g., Piro et al. 2000). The work of Kvamme
(2006) has provided clear avenues for this form of analysis, and comparative studies have been
quickly published in the USA (e.g., Johnson 2006). The Kvamme paper explores not only
descriptive graphical solutions (two-dimensional overlays, redgreenblue colour composites,
translucent overlays), but develops both discrete and continuous data integration as a rigorous
interpretive script. Statistical interrogation within a GIS environment included K-means cluster,
principal components and factor analysis, which are far from common in the analysis of
geophysical data sets. However, there is much more to be achieved in this area, as there is no
clear best approach, and an uneasy partnership exists between visually pleasing and statistically
different assessments. Additionally, some caution must be noted, because although the
theoretical benefits are obvious, considerable thought must go into the research design to
ensure that each data strand is appropriate for the study, both in interpretative capacity and
data quality/densitythe worry is that techniques may be collected like stamps, rather than
justified from an archaeological perspective (for a particular view on multi-technique surveys,
see Hesse 1999).
The fusion of data is a fundamental problem that exists beyond archaeological prospecting
and there are also significant theoretical obstacles to be cleared prior to true data fusion.
Additionally, some researchers have indicated the limits of GIS-based schemes. While GIS
provides a spatial context, there are limitations in registering volume data, and the inherent
three-dimensionality of GPR and electrical imaging/tomography data requires a different
approach to visualization and data fusion. Data integration and solid modelling are part of the
3-D cube approach that Watters (2006) has championed (Fig. 14). As data density increases
across the board and high-precision comparative data sets become more common, then
integration and visualization of survey data will become ever more important. As there is now
a trend to incorporate more than one type of detector on vehicle-drawn systems (Hill et al.
2004), developments within this field are likely to be rapid.
Other techniques
It is difficult to review the less frequently used techniques, but they are underused because they
are either too slow, costly or the application areas are tangential to the majority of archaeological
research. However, that is not to say that significant progress has been achieved. Gravity
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

330

C. Gaffney

Figure 13 Caesium Vapour Magnetometer data: a GIS-based archaeological interpretation map of enclosures within
a large, late Iron Age settlement at Roseldorf, Austria. The insert shows a reconstruction drawing based on similar
excavated enclosures, interpreted as a temple (image courtesy of W. Neubauersee Neubauer (2004) for further
details. Copyright 2004 John Wiley & Sons Limited. Reproduced with permission.)

University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

A review of geophysical techniques in archaeology

331

Figure 14 Data fusion and visualization using software (Amira) originally designed for medical imaging:
the image includes both visualization of data as well as interpreted elements in a micro-survey 10 10 m in extent
(image courtesy of M. Watters).

measurements have been used only sporadically, for identifying potentially significant voids
under buildings or use on greenfield sites (Linford 1998). Seismic studies have provided much
to the community by way of processing and visualization, but have had fewer applications
over the years. Whereas the earlier papers used refracted waves (Goulty and Hudson 1994),
the most recent have attempted to use ultra-shallow reflected waves to identify cavities
(Metwaly et al. 2005) and palaeochannels (Hildebrand et al. 2007). It is evident that this will
be a fertile research area in the future, and important papers can be found in related journals
such as Near Surface Geophysics, the Journal of Applied Geophysics and Archaeological
Prospection. Analysis of seismic data collected in the North Sea for oil and gas exploration
has revealed elements of the Mesolithic landscape (Fig. 15) on a scale that no terrestrial
technique has matched (Gaffney et al. 2007). On the smaller scale, induced polarization may
well become more frequently used, particularly in the search for buried trackways or in the
investigation of industrial sites, especially as some commercial electrical imaging systems
have time-domain measurements built in.
SOME CONCLUDING REMARKS

The review has highlighted some of the successes of ground-based geophysical survey over
the past 50 years, as well as indicating some current research themes that will be significant in
the next decade. Many of the techniques for terrestrial geophysicsinstrumentation, applications
and processinghave changed dramatically in the past decade. There is every indication that
this momentum will continue.
Some of the overarching themes are of greater interest to general archaeometrists. Real-time
GPS is now enabled on some instruments and the next generation will be fully linked to highprecision devices. Visualization will become increasingly important as data sets get larger and
more data dense, and the requirement to combine and interrogate disparate sources grows.
This will be linked to advanced processing that reveals more interpretable anomalies. It is
likely that this trend will be augmented by a return to a statistical analysis of data sets.
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

332

C. Gaffney

Figure 15 An analysis of seismic data originally collected for oil and gas exploration can reveal archaeological
landscapes on a scale previously unknown from geophysical techniques (image courtesy of V. Gaffney).

However, rather than simple anomaly recognition, this may be linked to texture, perhaps
indicating different types of archaeological activity as well as automatic anomaly identification
and interpretation. It is correct that the final trend should be linked to interpretation, as that is
the whole point of our discipline. As data sets grow, the recognition and interpretation of
small-scale anomalies becomes increasingly difficult and we must rise to that challenge. For
example, the work of Benech (2007) is a sophisticated analysis of social space based on
magnetometer data, and amply illustrates the technical, statistical and interpretation skills that
are required if the value of geophysical techniques to archaeology is to be fully recognized.
The geophysical toolbox is regularly opened for archaeological purposes and used throughout
the world. In terms of global applicability and scale of applications, from postholes to landscapes, geophysical techniques are highly desirable in many archaeological situations. In fact,
the take-up of magnetometry, earth resistance and GPR is such that, for some archaeologists,
geophysical techniques could be the most used science-based application in archaeology. While
it is acknowledged that for some a black box mentality exists, the science behind the machines
and the interpretation of data-rich environments that result from the endeavours of archaeological
prospection are still evolving as we strive to deliver new insights into the buried past.
ACKNOWLEDGEMENTS

I would like to thank the many colleagues who have helped in the production of this review.
Some have been named in the text, but many others remain anonymous. The latter are not
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

A review of geophysical techniques in archaeology

333

forgotten and I wish to acknowledge their inspiration in writing this review: without their hard
work, the archaeological world would be a duller place.

REFERENCES
Aitken, M. J., 1958, Magnetic prospecting. I. The Water Newton survey, Archaeometry, 1, 24 6.
Aitken, M. J., 1986, Proton magnetometer prospection: reminiscences of the first year, Prospezioni Archaeologiche,
10, 15 17.
Alldred, J. C., 1964, A fluxgate gradiometer for archaeological surveying, Archaeometry, 7, 14 19.
Aspinall, A., and Gaffney, C., 2001, The Schlumberger arraypotential and pitfalls in archaeological prospection,
Archaeological Prospection, 8, 199209.
Aspinall, A., and Lynam, J. T., 1970, An induced polarization instrument for the detection of near surface features,
Prospezioni Archeologiche, 5, 6775.
Aspinall, A., and Saunders, M. K., 2005, Experiments with the square array, Archaeological Prospection, 12, 115 29.
Aspinall, A., Gaffney, C., and Schmidt, A., in press, Magnetometry for archaeology, AltaMira Press, Walnut Creek, CA.
Astin, T., Ekardt, H., and Hay, S., 2007, Resistivity imaging survey of the Roman barrows at Bartlow, Cambridgeshire,
UK, Archaeological Prospection, 14, 2437.
Atkinson, R. J. C., 1953, Field archaeology, Methuen, London.
Aubrey, L., Benech, C., Marmet, E., and Hesse, A., 2001, Recent achievements and trends for research for geophysical
prospection of archaeological sites, Journal of Radioanalytical and Nuclear Chemistry, 247, 6218.
Bates, M. R., Bates, C. R., and Whittaker, J. E., 2007, Mixed method approaches to the investigation and mapping of
buried Quaternary deposits: examples from southern England, Archaeological Prospection, 14, 104 29.
Becker, H., 1995, From nanotesla to picoteslaa new window for magnetic prospecting in archaeology, Archaeological
Prospection, 2, 21728.
Becker, H., 2001, Duo- and quadro-sensor configuration for high speed/high resolution magnetic prospecting with
caesium magnetometer, in Magnetic prospecting in archaeological sites (eds. H. Becker and J. W. E. Fassbinder),
205, ICOMOS, Munich.
Benech, C., 2007, New approach to the study of city planning and domestic dwellings in the ancient Near East,
Archaeological Prospection, 14, 87103.
Benech, C., and Marmet, E., 1999, Optimum depth of investigation and conductivity response rejection of the different
electromagnetic devices measuring apparent magnetic susceptibility, Archaeological Prospection, 6, 31 45.
Berard, B. A., and Maillol, J. M., in press, Common- and multioffset ground penetrating radar study of a Roman villa,
Tourega, Portugal, Archaeological Prospection.
Berge, M., and Drahor, M. G., 2007, Electrical resistivity inversion modelling studies for commonly used arrays in
hyk (artificial hill) type of archaeological settlements, Studijne Zvesti, 41, 11216.
Bevan, B., 2001, An early geophysical survey at Williamsburg, USA, Archaeological Prospection, 7, 518.
Chapman, H., 2006, Landscape archaeology and GIS, Tempus, Stroud.
Chwala, A., Stolz, R., Ijsselsteijn, R., Schultze, V., Ukhansky, N., Meyer, H.-G., and Schler, T., 2001, SQUID gradiometers for archaeometry, Superconductor Science and Technology, 14, 111114.
Chwala, A., Ijsselsteijn, R., May, T., Oukhanski, N., Schler, T., Schultze, V., Stolz, R., and Meyer, H.-G., 2003,
Archaeometric prospection with high-TC SQUID gradiometers, IEEE Transactions on Applied Superconductivity,
13, 76770.
Clark, A., 1990, Seeing beneath the soil, Batsford, London.
Cole, M. A., Linford, N. T., Payne, A. P., and Linford, P. K., 1995, Soil magnetic susceptibility measurements and
their application to archaeological site investigation, in Science and site: evaluation and conservation (eds. J. Beavis
and K. Barker), 114 62, Bournemouth University, Bournemouth, UK.
Conyers, L. B., 2004, Ground-penetrating radar for archaeology, AltaMira Press, Walnut Creek, CA.
Conyers, L. B., and Goodman D., 1997, Ground-penetrating radar: an introduction for archaeologists, AltaMira
Press, Walnut Creek, CA.
Dabas, M., in press, Theory and practice of the new fast electrical imaging system ARP, in Geophysics for landscape
archaeology (eds. S. Campana and S. Piro), University of Siena.
Dabas, M., Blin, O., and Benard, C., 2005, Les nouvelles techniques de rsistivit lectrique employes dans la
prospection de grandes surfaces en archologie, Les Nouvelles de lArchologie, 101, 24 32.
Dabas, M., Hesse, A., and Tabbagh, J., 2000, Experimental resistivity survey at Wroxeter archaeological site with a
fast and light recording device, Archaeological Prospection, 7, 10718.
University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

334

C. Gaffney

Dabas, M., Decriaud, J. P., Ducomet, G., Hesse, A., Mounir, A., and Tabbagh, A., 1994, Continuous recording of
resistivity with towed arrays for systematic mapping of buried structures at shallow depth, Revue dArchomtrie,
18, 1319.
Dalan, R., 2008, A review of the role of magnetic susceptibility in archaeogeophysical studies in the United States:
recent developments and prospects, Archaeological Prospection, 15, 131.
David, A., Cole, M., Horsley, T., Linford, N., Linford, P., and Martin, L., 2004, A rival to Stonehenge? Geophysical
survey at Stanton Drew, England, Antiquity, 78, 34158.
Fassbinder, J. W. E., and Stanjek, H., 1993, Occurrence of bacterial magnetite in soils from archaeological sites,
Archaeologia Polona, 31, 11728.
Fassbinder, J. W. E., Stanjek, H., and Vali, H., 1990, Occurrence of magnetic bacteria in soil, Nature, 343, 1613.
Fowler, P. J., 1959, Magnetic prospecting: an archaeological note about Madmarston, Archaeometry, 2, 379.
Gaffney, C., and Gaffney, V., 2006, No further territorial demands: on the importance of scale and visualization
within archaeological remote sensing, paper presented at From Artefacts to Anomalies: Papers inspired by the
contribution of Arnold Aspinall, University of Bradford, 12 December 2006; http://www.brad.ac.uk/archsci/
conferences/aspinall/
Gaffney, C., and Gater, J., 2003, Revealing the buried past: geophysics for archaeologists, Tempus, Stroud.
Gaffney, V., Thomson, K., and Fitch, S. (eds.), 2007, Mapping Doggerland: the Mesolithic landscapes of the southern
North Sea, Archaeopress, Oxford.
Gaffney, C. F., Gater, J. A., Linford, P., Gaffney, V. L., and White, R., 2000, Large-scale systematic fluxgate
gradiometry at the roman city of Wroxeter, Archaeological Prospection, 7, 8199.
Goulty, N. R., and Hudson, A. L., 1994, Completion of the seismic refraction survey to locate the vallum at Vindobala,
Hadrians Wall, Archaeometry, 36, 32735.
Griffiths, D. H., and Barker, R. D., 1994, Electrical imaging in archaeology, Journal of Archaeological Science, 21,
153 8.
Hesse, A., 1966, Prospections gophysiques fiable profondeur: applications larchologie, Dunod, Paris.
Hesse, A., 1999, Multi-parametric survey for archaeology: How and why, or how and why not? Journal of Applied
Geophysics, 41, 15768.
Hildebrand, J. A., Wiggins, S. M., Driver, J. L., and Waters, M. R., 2007, Rapid seismic reflection imaging at the
Clovis period Gault site in central Texas, Archaeological Prospection, 14, 245 60.
Hill, I., Grossey, T., and Leech, C., 2004, High-resolution multi-sensor geophysical surveys for near surface applications can be rapid and cost-effective, The Leading Edge, 684 9.
Hoskins, W. G., 1976, The making of the English landscape, 2nd edn, Penguin, Middlesex (first published 1956).
Johnson, J. K. (ed.), 2006, Remote sensing in archaeology: an explicitly North American perspective, University of
Alabama Press, Tuscaloosa, AL.
Kamei, H., Marukawa, Y., Kudo, H., Nishimura, Y., and Nakai, M., 2000, Geophysical survey of HiruiOtsuka
Mounded tomb in Ogaki, Japan, Archaeological Prospection, 7, 22530.
Keay, S., and Millett, M., 2000, Falerii Novi: a new survey of the walled area, Papers of the British School at Rome,
68, 193.
Kelly, M. A., Dale, P., and Haigh, J. B. H., 1984, A microcomputer system for data logging in geophysical surveying,
Archaeometry, 26, 18391.
Kvamme, K. L., 2003, Geophysical surveys as landscape archaeology, American Antiquity, 63, 43557.
Kvamme, K. L., 2006, Integrating multidimensional archaeological data, Archaeological Prospection, 13, 5772.
Leckebusch, J., 2003, Ground-penetrating radar: a modern three-dimensional prospection method, Archaeological
Prospection, 10, 21340.
Leckebusch, J., 2005, Precision real-time positioning for fast geophysical prospection, Archaeological Prospection,
12, 199202.
Leckebusch, J., 2007, Pull-up/pull-down corrections for ground-penetrating radar data, Archaeological Prospection,
14, 1425.
Linford, N. T., 1998, Geophysical survey at Boden Vean, Cornwall, including an assessment of the microgravity
technique for the location of suspected archaeological void features, Archaeometry, 40, 187216.
Linford, N. T., 2006, The application of geophysical methods to archaeological prospection, Reports on Progress in
Physics, 69, 220557.
Linford, N., Linford, P., Martin, L., and Payne, A., 2007, Recent results from the English Heritage caesium
magnetometer system in comparison with recent fluxgate gradiometers, Archaeological Prospection, 14, 151
66.
Linington, R. E., 1964, The use of simplified anomalies in magnetic surveying, Archaeometry, 7, 313.

University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

A review of geophysical techniques in archaeology

335

Loke, M. H., and Barker, R. D., 1996, Rapid least-squares inversion of apparent resistivity pseudosections by a
quasi-Newton method, Geophysical Prospecting, 44, 13152.
Metwaly, M., Green, A. G., Horstmeyer, H., Maurer, H., Abbas, A. M., and Hassaneen, A. Gh., 2005, Combined
seismic tomographic and ultrashallow seismic reflection study of an Early Dynastic mastaba, Saqqara, Egypt,
Archaeological Prospection, 12, 24556.
Neubauer, W., 2001, Magnetische Prospektion in der Archologie, Mitteilungen der Prhistorischen Kommission,
Wien.
Neubauer, W., 2004, GIS in archaeologythe interface between prospection and excavation, Archaeological Prospection,
11, 159 66.
Neubauer, W., and Eder-Hinterleitner, A., 1997, Resistivity and magnetics of the Roman town Carnuntum, Austria:
an example of combined interpretation of prospection data, Archaeological Prospection, 4, 179 89.
Noel, M., and Xu, B., 1991, Archaeological investigation by electrical resistivity tomography: a preliminary study,
Geophysical Journal International, 107, 95102.
Nuzzo, L., 2005, Identification and removal of above-ground spurious signals in GPR archaeological prospecting,
Archaeological Prospection, 12, 93103.
Panissod, C., Dabas, M., Florsch, N., Hesse, A., Jolivet, A., Tabbagh, A., and Tabbagh, J., 1998, Archaeological prospecting using electric and electrostatic mobile arrays, Archaeological Prospection, 5, 239 51.
Papadopoulos, N. G., Tsourlos, P., Tsokas, G. N., and Sarris, A., 2006, Two-dimensional and three-dimensional resistivity imaging in archaeological site investigation, Archaeological Prospection, 13, 163 81.
Perssona, K., and Olofsson, B., 2004, Inside a mound: applied geophysics in archaeological prospecting at the Kings
Mounds, Gamla Uppsala, Sweden, Journal of Archaeological Science, 31, 551 62.
Pipan, M., Baradello, L., Forte, E., Prizzon, A., and Finetti, I., 1999, 2-D and 3-D processing and interpretation of
multi-fold ground penetrating radar data: a case history from an archaeological site, Journal of Applied Geophysics,
41, 27192.
Piro, S., Mauriello, P., and Cammarano, F., 2000, Quantitative integration of geophysical methods for archaeological
prospection, Archaeological Prospection, 7, 20313.
Powlesland, D., Lyall, J., Hopkinson, G., Donoghue, D., Beck, M., Harte, A., and Stott, D., 2006, Beneath the sand
remote sensing, archaeology, aggregates and sustainability: a case study from Heslerton, the Vale of Pickering,
North Yorkshire, UK, Archaeological Prospection, 13, 2919.
Ralph, E. K., 1964, Comparison of a proton and rubidium magnetometer for archaeological prospecting, Archaeometry,
7, 20 7.
Schultze, V., Chwala, A., Stolz, R., Schulz, M., Linzen, S., Meyer, H.-G., and Schler, T., 2007, A superconducting quantum interference device system for geomagnetic archaeometry, Archaeological Prospection, 14, 226
9.
Schultze, V., Linzen, S., Schler, T., Chwala, A., Stolz, R., Schulz, M., and Meyer, H.-G., in press, Fast and sensitive
geomagnetic archaeometry of large areas using SQUIDsthe measurement system and its application to the
Niederzimmern Neolithic double-ring ditch exploration, Archaeological Prospection.
Scollar, I., 1968, A program package for the interpretation of magnetometer data, Prospezioni Archeologice, 3, 918.
Scollar, I., Tabbagh, A., Hesse, A., and Herzog, I., 1990, Archaeological prospecting and remote sensing, Cambridge
University Press, Cambridge.
Somers, L., Linford, N., Penn, W., David, A., Urry, L., and Walker, R., 2005, Fixed frequency radio wave imaging of
subsurface archaeological features: a minimally invasive technique for studying archaeological sites, Archaeometry,
47, 15973.
Spoerry, P. (ed.), 1992, Geoprospection in the archaeological landscape, Oxbow Monograph 18, Oxbow Books,
Oxford.
Sutherland, T. L., and Schmidt, A., 2003, Towton 1461: an integrated approach to battlefield archaeology, Landscapes, 4,
1525.
Tabbagh, A., 1986a, Applications and advantages of the slingram electromagnetic method for archaeological
prospecting, Geophysics, 51, 57684.
Tabbagh, A., 1986b, What is the best coil orientation in the slingram electromagnetic prospecting method? Archaeometry,
28, 18596.
Tite, M. S., 1966, Magnetic prospecting near to the geomagnetic equator, Archaeometry, 9, 2431.
Tonkov, N., and Loke, M. H., 2006, A resistivity survey of a burial mound in the Valley of the Thracian Kings,
Archaeological Prospection, 13, 12936.
Tsokas, G. N., Tsourlos, P. I., Vargemezis, G., and Novack, M., in press, Non-destructive ERT for indoor investigation:
the case of Kapnikarea church in Athens, Archaeological Prospection.

University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

336

C. Gaffney

Tsokas, G. N., Tsourlos, P., Papadopoulos, N., Manidaki, V., Ioannidou, M., and Sarris, A., 2006, Non destructive
ERT survey at the South Wall of the Akropolis of Athens, in Extended abstracts of Near Surface 2006, B040,
Helsinki, Finland.
Venter, M. L., Thompson, V. D., Reynolds, M. D., and Waggoner, J. C. Jr, 2006, Integrating shallow geophysical
survey: archaeological investigations at Totogal in the Sierra de los Tuxtlas, Veracruz, Mexico, Journal of
Archaeological Science, 33, 76777.
Vickers, R. S., and Dolphin, L. T., 1975, A Communication on an archaeological radar experiment at Chaco Canyon,
New Mexico, MASCA Newsletter, 11, 68.
Walker, R., 2000, Multiplexed resistivity survey at the Roman town of Wroxeter, Archaeological Prospection, 7, 119
32.
Walker, R., and Linford, P., 2006, Resistance and magnetic surveying with the MSP40 Mobile Sensor Platform at
Kelmarsh Hall, ISAP News, 9, 3 4.
Walker, R., Gaffney, C., Gater, J., and Wood, E., 2005, Fluxgate gradiometry and square array resistance survey at
Drumlanrig, Dumfries and Galloway, Scotland, Archaeological Prospection, 12, 131 6.
Walton, D., 1977, Archaeomagnetic intensity measurements using a SQUID magnetometer, Archaeometry, 19, 192
200.
Watters, M. S., 2006, Geovisualization: an example from the Catholme Ceremonial Complex, Archaeological
Prospection, 13, 28290.
Witten, A. J., Thomas, E., Levy, T. E., Adams, R. B., and Won, I. J., 2000, Geophysical surveys in the Jebel Hamrat
Fidan, Jordan, Geoarchaeology, 15, 13550.
Wynn, J. C., 1986, Review of geophysical methods used in archaeology, Geoarchaeology, 1, 24557.

University of Oxford, 2008, Archaeometry 50, 2 (2008) 313336

También podría gustarte