Está en la página 1de 17

This article was downloaded by: [82.24.31.

176]
On: 11 December 2014, At: 06:48
Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Structure and Infrastructure Engineering:


Maintenance, Management, Life-Cycle Design and
Performance
Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/nsie20

Uplift of elastomeric bearings in isolated bridges


subjected to longitudinal seismic excitations
Stergios A. Mitoulis

Department of Civil and Environmental Engineering, Faculty of Engineering and Physical


Sciences, University of Surrey, Guildford, Surrey GU2 7XH, UK
Published online: 24 Nov 2014.

To cite this article: Stergios A. Mitoulis (2014): Uplift of elastomeric bearings in isolated bridges subjected to longitudinal
seismic excitations, Structure and Infrastructure Engineering: Maintenance, Management, Life-Cycle Design and Performance,
DOI: 10.1080/15732479.2014.983527
To link to this article: http://dx.doi.org/10.1080/15732479.2014.983527

PLEASE SCROLL DOWN FOR ARTICLE


Taylor & Francis makes every effort to ensure the accuracy of all the information (the Content) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.
This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions

Structure and Infrastructure Engineering, 2014


http://dx.doi.org/10.1080/15732479.2014.983527

Uplift of elastomeric bearings in isolated bridges subjected to longitudinal seismic excitations


Stergios A. Mitoulis*
Department of Civil and Environmental Engineering, Faculty of Engineering and Physical Sciences, University of Surrey, Guildford,
Surrey GU2 7XH, UK

Downloaded by [82.24.31.176] at 06:48 11 December 2014

(Received 12 February 2014; final version received 21 August 2014; accepted 15 September 2014)
Bearings are used to isolate bridge substructures from the lateral forces induced by creep, shrinkage and seismic
displacements. They are set in one or two support lines parallel to the transverse axis of the pier cap and are typically
anchored to the deck and to the pier cap. This detailing makes them susceptible to possible tensile loading. During an
earthquake, the longitudinal displacements of the deck induce rotations to the pier caps about a transverse axis, which in turn
cause tensile (uplift) and compressive displacements to the bearings. Tensile displacements of bearings, due to the pier
rotations, have not been addressed before and questions about the severity of this uplift effect arise, because tensile loading
of bearings is strongly related to elastomer cavitation and ruptures. An extended parametric study revealed that bearing
uplift may occur in isolated bridges, while uplift effect is more critical for the bearings on shorter piers. Tensile
displacements of bearings were found to be significantly increased when the isolators were eccentrically placed with respect
to the axis of the pier and when flexible isolators were used for the isolation of the bridge. The results of this study cannot be
generalised as bridge response is strongly case-dependent and the approach has limitations, which are related to the
modelling approach and to the fact that emphasis was placed on the longitudinal response of bridges.
Keywords: bridges; bearings; uplift; longitudinal displacement; rotation; pier cap

1. Introduction
Seismic isolation has been used extensively during the last
years in bridges. An isolation system placed between the
bridge superstructure and its supporting substructure
lengthens the fundamental period of the bridge structure
such that the bridge does not respond to the most damaging
energy content of the earthquake input (Kunde & Jangid,
2003), while offering damping to the bridge. The
earthquake-resisting system (ERS) of isolated bridges
relies heavily on the isolators. For this reason, current code
design provisions for isolated bridges (American Association of State Highway and Transportation Officials
[AASHTO], 2010; Eurocode 8-2, 2005; Japan Road
Association [JRA], 1997, 2002) are quite demanding,
especially on issues related to the reliability of the
isolation system, to prevent potential deck unseating that
jeopardise the integrity of the bridge (Imbsen, 2007;
National Institute for Standards and Technology [NIST],
1996), while EN 1337-3 (2005) section 8.2.1.2.7 and BS
EN 15129 (2009) require that tensile stresses should not be
greater than 2G, where G is the shear modulus measured at
100% shear strain, to avoid cavitation.
According to reconnaissance reports, among damaged
bridges, bearing uplift and ruptures and the consequent span
unseating as well as pounding effects were the most common
failure modes after the 1971 San Fernando earthquake, the

*Email: s.mitoulis@surrey.ac.uk
q 2014 Taylor & Francis

1989 Loma Prieta earthquake (Priestley, Seible, & Calvi,


1996), the 1994 Northridge earthquake (Earthquake
Engineering Research Institute [EERI], 1995a), the 1995
Kobe earthquake (EERI, 1995b), the 1999 Chi Chi
earthquake (EERI, 2001) and the 2011 Tohoku earthquake
(Akiyama, Frangopol, & Mizuno, 2014; Buckle, Yen,
Marsh, & Monzon, 2012; EERI, 2011; Kawashima, 2012;
Kitahara, Kajita, & Kitane, 2012). The span unseating and
the bearing uplift dread are evident throughout most bridge
design codes (California Department of Transportation
[CalTrans], 1999; EN 1998-2, 2005; JRA, 2002).
Damaged bearings, following severe earthquakes, are
a quite frequent bridge failure mode as observed by Chang,
Chang, Tsai, and Sung (2000). According to Chu-Chieh,
Hung, Liu, and Chai (2010), the uplift failure of the
bearings was one of the main reasons for the collapse of
Baihwa Bridge. Similarly, the Dong Feng bridge girders
were dislodged and fallen due to the failure of the
supporting bearings, while Ji-lu bridge suffered similar
span unseating (Chang et al., 2000). Ruptures of bearings
were also observed in the last Tohoku earthquake. Nearly
200 highway bridges and numerous rail bridges were
damaged during the earthquake by effects including span
unseating and ruptured bearings (EERI, 2011). The
Sendai-Tohbu Viaduct revealed damages and ruptures at
four of its pier elastomeric bearings (Public Works

Downloaded by [82.24.31.176] at 06:48 11 December 2014

S.A. Mitoulis

Research Institute [PWRI], 2011) and some of girders


were separated from the bearings.
Kawashima (2012) attributed these damages primarily
to inadequate design and fabrication of the elastomeric
bearings and secondly to the interaction between adjacent
decks. Buckle et al. (2012) suggested that the hammerhead
piers rotated about a longitudinal axis twisting the
superstructure about the same axis, indicating at potential
bearing unseating effects due to the seismic movements of
the deck in its transverse direction. Katsaras, Panagiotakos, and Kolias (2009) studied the effect of bearing
uplifting when the bridge is subjected to seismic loading in
the transverse direction. The uplift of anchored bearings
was found to be severe for the bearings on the abutments.
The rupture of bearings anchored to piers, as if in direct
tension (Buckle et al., 2012) during the 2011 Tohoku
earthquake, was confirmed for the first time since isolation
bearings were adopted according to Kitahara et al. (2012).
It seems that bridges that had the bearings neither anchored
nor glued to the girder (Hamzeh, Tassoulas, & Becker,
1998) or the pier caps have suffered slippage or walking
out of place according to McDonald, Heymsfield, and
Avent (2000) and Aria and Akbari (2013), but they avoided
ruptures under tensile loading. Hence, uplift of anchored
bearings is an undesirable loading condition, hence design
equations for bearings subjected to tensile loads are not
provided as stated by Constantinou, Kalpakidis, Filiatrault,
and Ecker Lay (2011).
Current codes such as AASHTO (2013) Methods A
and B design of elastomeric bearing require that no
bearing uplift occurs due to the rotation of the bearing
under seismic loads. According to NCHRP 596 (Stanton
et al., 2008), the original reason to include the no uplift
provision in the standard was to prevent internal rupture of
the bearing. Similarly, Eurocode 8 (Eurocode 8-2 section
6.6.3.2, EN 1998-2, 2005) requires that no uplift of
isolators occurs under the design seismic combination.
Accordingly, Kelly (2003) and Kelly and Takhirov (2007)
found that bearing uplift is not detrimental to bearing
integrity. As it is stated in these studies, the interaction of
shear- and tension-induced stresses is a factor that
increases the resistance of the bearing; however, the
(a)

onset of cavitation in elastomer occurs at low axial strains,


in the order of 1023. Indeed, Yang, Liu, He, and Feng
(2010) found that elastomeric bearing response under
tensile loading is expected to be nonlinear. Bearings were
found to yield at tensile strains 3%, while post-elastic
stiffness is 6% of the initial elastic one.
Iwabe, Takayama, Kani, and Wada (2000) found that
even if 100% tensile strain was induced, rubber bearings
did not rupture. Artieda and Whittaker (2010) found that
bridges may exhibit unseating even when the seismic
actions are evenly distributed or under serviceability
loading, namely for quite small deck displacements.
Indeed, rupture of anchored bearings due to uplift
displacements is a possible failure mode according to
Dorfmann and Burtscher (2000). Hence, questions about
the magnitude of bearing uplift displacements in bridges
subjected to longitudinal seismic loadings arise (Mitoulis,
Muhr, & Ahmadi, 2014).
This paper sheds light on an unknown uplift
mechanism of elastomeric bearings that are placed
eccentrically with respect to the axis of the pier. Current
code provisions such as AASHTO (2010) figure 7.1-1,
shown in Figure 1(a), and Eurocode 8 Part 2 at the key
figure of section 7.5.4 (2005), shown in Figure 1(b), do not
provide a bearing-pier model that accounts for this uplift
mechanism, nor for the nonlinear response of elastomeric
bearings under tensile loading. The two figures also show
that different choices are given for the boundary
conditions assumed for the bearings during design,
which suggest that this issue deserves attention. The
magnitude of bearing uplift displacements is estimated
through an extended parametric study. Emphasis is being
placed on the nonlinear response of elastomeric bearing in
its vertical direction.
2.

Description of the uplift effect of bearings

2.1. Development of the uplift of bearings


The description of how tensile stresses are developed in
the bearings is provided in this section. Figure 2 presents
the initial stage of compressive deformation of bearings
due to the weight of the deck Wd that corresponds to the
Superstructure

(b)

Isolator
(one or more bearings)

keff

did
dbi,d

deck
Fi

Substructure
bearing

ksub
Hi

pier
di

dsub
d

Figure 1.

(a) AASHTO (2010) figure 7.1-1 and Eurocode 8 Part 2 (2005) key figure of section 7.5.4.

Structure and Infrastructure Engineering

deformation of both lines of bearings is almost the same


aver
(vaver
b1 ;0 < vb2 ;0 ).
Three indices were used for the vertical deformations
of the bearings vb, as shown in Figure 2: the first lower
index indicates whether bearings are placed in line 1 or 2.
The second lower index refers to the loading phase. The
superscript of vb indicates the vertical deformation of the
bearing left (l) or right (r) side, while aver denotes the
average vertical deformation of the bearing. Hence, vlb1 ;0
is the vertical deformation on the left side of the bearings
placed at line 1 (left side of the pier cap) during Phase 0.
Figures 2 and 3 show by dashed lines the initial
(unloaded) bearing geometry and the potential deformed
shapes of the isolators while longitudinal movement of the
deck occurs. The analyses did not account for the
movements of the bearings due to thermal, creep and
shrinkage effects. These movements are expected to
influence the response of the bearings when the length of
the bridge is large.

deck weight Wd

line 1

line 2

Vb1,0

Vb1,0

Vb2,0

r
Vb1,0 + Vb1,0
(compr.)
2

aver

Vb1,0 =

Vb2,0

aver

Vb2,0 =

Vb2,0 + Vb2,0
(compr.)
2

Downloaded by [82.24.31.176] at 06:48 11 December 2014

Figure 2. The bearings are under compression due to the weight


of the deck (Phase 0).

permanent, additional permanent and 20% of the live


(variable) loadings. This phase is denoted as Phase 0.
The bearing on the left line support represents the first line
of the bearings (line 1), while the bearing on the right line
support represents the second line (line 2). All the
bearings are under compressive load. The initial vertical

d1

(a)
ub,1

ub,1

(b)

d2

up,1

vb1,1

ub,2

ub,1

ub,1

vb2,1

vb1,1

ub,2

vb2,1

vb1,1=
aver

vb1,2

vb1,1 + v
2

pier cap's vertical


movement vp 1 e

aver

vb2,1 =

Detail
pier cap's upper face
1

vb2,1 + vb2,1
(compr.)
2
1

vp

aver

vb2,2=

(d)

d4

u b,4

u b,3

u p,3

v b1,3
r

u b,4

v b1,4

u b,3

u b,3

(compr.)

Bearings at line 2 are under slight tensile


deformation (Phase 3).

aver

r
vb2,3
aver
(tension)
vb2,3 =
2

u b,4

v b2,4
r

v b2,4

v b2,3
vb1,3 + v
2

u b,4

vb2,3 =0

u p,4
v b1,4

vb1,3

aver

vb2,2
(compr.)
2

Compression of the bearings at line 1 is further


increased. Bearings at line 2 are under slight
compressive deformation (Phase 2).

d3

vb1,3 =

r
vb2,2

r
vb1,2 + vb1,2
(compr.)
2
2

Compression of the bearings at line 1 is


increased. Compressive displacements of
bearings at line 2 are reduced (Phase 1).

(c)

aver

vb1,2 =

ub,2

eccentricity e
pier caps upper face
rotates clockwise

r
b1,3

ub,2
vb2,2

(compr.)

u b,3

up,2

vb1,2

r
b1,1

vb1,4 =

r
vb1,4 + vb1,4
(compr.)
2

4
r

aver

vb2,4 =

vb2,4 + vb2,4
(tension)
2

Maximum compression of the bearings at line


1. Bearings at line 2 are under max tensile
deformation (Phase 4).

Figure 3. Description of progressive bearings shear, rotational and vertical (compressive and tensile) deformations due to the
longitudinal displacement (ud) of the deck and the rotation (w) of the pier cap about a transverse axis (indicative).

Downloaded by [82.24.31.176] at 06:48 11 December 2014

S.A. Mitoulis

The following description considered that the longitudinal, transverse and vertical axes are x, y and z,
respectively. When the deck displaces in along the
longitudinal direction x, either due to seismic movements
or due to thermal movements, creep and shrinkage
(Mitoulis, 2012), then the following sequential phases of
bearings deformation are anticipated: during Phase 1,
which is illustrated in Figure 3(a), the deck is displaced by
ud1 . The pier cap exhibits longitudinal (along x)
displacement up,1 and rotation w1 about its transverse
axis (y), due to the flexural deflection of the pier and the
rotation of the foundation.
This pier-cap rotation (w1) imposes both rotations and
vertical (along z) deformations to the bearings. The first
line of the bearings (on the left side of the pier cap, i.e. line
1 at Figure 3(a)) is likely to receive higher compression
aver
loads than in Phase 0 (i.e. vaver
b1 ;1 < vb1 ;1 ) as pier cap moves
upwards by vp, shown in the detail of Figure 3(a). The
opposite occurs at the right side of the pier cap. The pier
caps upper face moves downwards (along z) and pulls the
bearings downwards. Hence, the compressive deformation
of bearings at line 2 is expected to be reduced, i.e.
aver
vaver
b2 ;1 < vb2 ;0 . Apart from the shear (along x) and vertical
(along z) displacements of the bearings, the isolators are
also rotated about a transverse axis (y). The rotation of the
bearings induces differential vertical displacements at the
left vlb1 ;1 and right sides vrb1 ;1 of the isolators, shown in
Figure 3(a).
Progressively, the deck longitudinal displacements
increase ud1 , ud2 , ud3 , ud4 at Phases 1 4 correspondingly, as shown in Figure 3(a) (d). Hence, the pier-cap
rotation increases, i.e. w1 , w2 , w3 , w4. Likewise, the
bearings are sheared by gradually increased shear strains,
1s, and rotations, w. The bearings at line 1 are increasingly
compressed. Hence, their compressive deformation is
aver
aver
aver
expected to be increased vaver
b1 ;0 , vb1 ;1 , vb1 ;2 , vb1 ;3
aver
, vb1 ;4 . The bearings at line 2 are likely to gradually
reduce their compressive stresses. During Phase 2, which
is illustrated in Figure 3(b), the bearings at line 2 exhibit a
slight compressive deformation. During Phase 3, shown in
Figure 3(c), the bearings at line 2 are under a slight tensile
deformation, i.e. vrb2 ;3 . 0 at the right edge, while their left
side has almost zero vertical deformation (vlb2 ;3 < 0).
Under the maximum longitudinal displacement of the deck
ud4 (Phase 4), the bearings at line 2 are under tensile
loading. Both sides of the isolators exhibit tensile
displacements, namely both vlb2 ;4 and vrb2 ;4 are uplift
displacements.
2.2.

Equilibrium of actions during bearing uplift

The equilibrium of actions is derived with respect to the


centre of gravity (CG) of the pier cap. An approximate
equation is illustrated indicating the uplift mechanism of
bearings. Figure 4 shows the actions of the deck, the

Sad
Vd=Wd g

deck weight Wd

N.A.
hd

deck
Mb2
Nb2
Vb2
line 2
Mb2
Nb2
Vb2

Mb1
Nb1
Vb1
bearings line 1
Mb1
Nb1
Vb1
e e

pier cap

hb

hp

C. G.

Mp
Np

Vp
pier
reference
fibre

Figure 4. The actions of the deck, the bearings, the pier cap and
the pier when the deck displaces in the longitudinal direction
(along x).

bearings, the pier cap and the pier when the deck
induces a vertical load Wd and a horizontal load Vd to
the bearings. The case of an isolated bridge with two
lines of bearings supporting the I-beams of the
superstructure spans is considered. This eccentricity e
of the bearings is the longitudinal distance of the
bearings with respect to the CG of the pier cap. The pier
cap has a depth equal to hp.
In Figure 4, Wd is the weight of the deck; Vd is the
longitudinal (along x) seismic loading of the deck, i.e.
Vd WdSad/g, where Sad is the deck acceleration and g the
acceleration of gravity. M bi , N bi and V bi are the total
bending moment, axial load and shear load of the bearings
at line i, respectively. Line i is parallel to the transverse
direction (y) of the pier cap and consists of n bearings
placed in line. Mp, Np and Vp are the bending moment, the
axial load and the shear of the pier cap and the pier,
respectively. Hence, the total vertical load N bi of n
bearings placed at line i is expressed as follows:
N bi

n
X

N bij :

j1

If N bi , 0 then bearings of line i are under axial


compression, if N bi 0 then the bearings are unloaded,
whereas if N bi . 0 then the bearings of line i are under
axial tension, i.e. they exhibit uplift. The total axial load of
the bearings N bi is considered to be zero when the average
vertical displacement of the bearings vaver
bi , shown in
Figure 3(a) (d), is zero.
Assuming that (1) the bearings at lines 1 and 2 have
equal shear deformations, i.e. V b1 V b2 V b , hence
Vp 2Vb, and (2) the two lines of bearings have the same
rotations,1 namely equal to the relative pier deck rotation

Structure and Infrastructure Engineering


Table 1.

Successful prediction of bearing lift off by equilibrium equation.


(3) N b2
(kN)

(4) e
(m)

(5) Sad
(m/s2)

(6) h
(m)

(7) Equation (4),


W d ag =ghp 2 e

(8) Mpa (kN m)


(NLDTHA)

(9) %
Discrepancy

0.00 7141.9

1.50

1.91

1.0

9322.3

9057.7

1.50

3.95

1.0

7686.8

7248.0

Bearing line
that is uplifted

(1) Wd
(kN)

1 (Left)

7141.9

2 (Right)

7004.88 7004.90

(2) N b1
(kN)

0.00

Mp is dependent upon hd, hb and hp and the partial fixity of the pier by the bearings.

that has been described through Figure 3(a) (d), and


hence the same bending moments, i.e. M b1 M b2 M b ,
the equilibrium of actions with respect to the CG of the
pier cap yields
N b1 2 N b2 e V p hp 2M b 2 M p 0:

Downloaded by [82.24.31.176] at 06:48 11 December 2014

Equation (2) is the equilibrium of actions of the pier cap. If


N bi 0 then uplift originates at bearings places at line i.
If N b2 0, then Equation (2) yields
N b1 e V p hp 2M b 2 M p 0:

The bearings at line 1 receive a compressive load equal to


Wd, since N b2 0. If the bending moments of the bearings
are neglected (M b M b1 M b2 0) and replacing
V d W d Sad =g, Equation (3) yields


Sad
hp 2 e M p :
W d
g

Equation (4) is the equilibrium of actions of the pier cap at


the onset of the bearings uplift. The equation corresponds
to the uplift of the bearings at line 2 (i 2). Furthermore,
if Equation (2) is solved with respect to N b2 and consider
(M b M b1 M b2 0), the equation yields
N b2

N b1 e V p hp 2 M p
:
e

Based on Equation (5), the bearings of the second line of


support are under tension when N b2 . 0, i.e. when
N b1 e V p hp 2 M p . 0:

If the term N b1 is replaced by Np, i.e. the axial load of the


pier, then the equation yields the so-called uplift parameter
(UP) as follows:
UP N p e V p hp 2 M p . 0;

The analytical validation of the approximate


Equation (4) through nonlinear dynamic time history
analyses (NLDTHAs) showed that Equation (4) predicts
the onset of bearing uplift successfully. The maximum
discrepancy between the results obtained by Equation
(4) and the corresponding results of the NLDTHA was
up to 6%, as shown for pier 4 in Table 1. It is also
worth noting that, based on Table 1, the first uplift
takes place at a deck acceleration of around 1.91 m/s2,
i.e. 0.19 g.
The observed discrepancies, shown in Table 1, were
found to be due to the assumptions of the approximate
equilibrium. More specifically, the equilibrium did not
take into account the bending moments of the bearings Mb.
Based on further analyses, using refined finite element
models (FEMs), it was found that the bending moments of
the bearings, Mb, were negligible when compared with the
bending moments of the piers Mp, i.e. the Mb is up to 3.5%
of the Mp. Also, the approximate model did not consider
the damping forces of the isolators. However, the bearings
of the as-built bridge were low damping bearings.
In addition, the onset of the tensile loading of the bearings
was found to occur when the isolators were sheared under
the large longitudinal displacements of the deck, i.e. when
the longitudinal velocities of the deck were small, thus
causing negligible damping forces on the bearings.

when UP , 0, the bearings are under compression, while


for UP $ 0 the uplift of bearings is expected. The UP can
be used by Bridge Engineers to assess the potential of the
uplift of bearings by applying simplified analyses, such as
the fundamental mode method in the longitudinal direction
of the bridge.

3. Description of the benchmark bridge and


parametric study
3.1. The benchmark bridge
The uplift mechanism of the bearings was studied through
a parametric study. The study utilised a typical isolated
bridge with prestressed I-beam girders, which are common
in southern Europe. The bridge, which was recently built
along the major Egnatia Highway at Northern Greece, is
illustrated in Figure 5. It has a total length equal to
148.9 m. The two end spans have a length equal to
29.45 m, while the three intermediate spans are 30.0 m
long. The deck is supported on the abutments and on the
piers through 5 and 10 low damping rubber bearings,
respectively. The dimension of the bearings at piers 1 3
are 400 126(66) (diameter 400 mm, total thickness of
the bearing 126 mm and total thickness of the elastomer
66 mm, i.e. six layers of 11 mm), while bearings with

Downloaded by [82.24.31.176] at 06:48 11 December 2014

Figure 5.

S.A. Mitoulis

The geometry of the BB: longitudinal section of the bridge, cross sections of the deck, the pier and the foundation.

dimensions 450 186(110) (10 layers of elastomer with


thickness 11 mm each) were used for the support of the
deck at the abutments and at pier 4.
The deck consists of five prestressed and precast Ibeams, precast slabs and cast-in-situ part of the slab. The
transverse width of the deck is 13.45 m and the moments of
inertia of its cross section are 3.17 and 128.31 m4 about the
transverse y and the vertical z axes, respectively. The piers
are circular with a diameter equal to 2.50 m and heights 16,
31, 28 and 13 m. The bridge is founded on ground type B
according to Eurocode 8-1 (EN 1998-1, 2005), through 3
3 pile groups. The piles have a diameter 0.80 m and
length 15 m at piers 1 and 2, while the piles at piers 3 and 4
are 5 m long. The design ground acceleration was equal to
0.16 g. The importance factor adopted was equal to
gI 1.0 according to Eurocode 8-2 (EN 1998-2, 2005).
The reinforcement ratios of the piers were found to be
equal to the minimum code requirements (Eurocode 2-1-1,
2004), i.e. the longitudinal reinforcement ratio is
rmin 1%.
3.2. Modelling and parametric study
Different bridge models were analysed to assess the
magnitude and the severity of the bearing uplift
displacements, when the bridge is subjected to longitudinal seismic excitations. Emphasis was placed on the
modelling of the vertical response of the bearings, namely
the tensile and compressive stiffness of the isolators, and
on the eccentricity e of the bearings with respect to the axis
of the pier, which is illustrated in Figures 3(a) and 4.
Further analyses were carried out to identify the influence
of specific bridge design parameters on the bearing uplift
displacements. The parameters were the following: (1) the
bearing modelling, by incorporating both linear and
nonlinear vertical stiffness models and different shear and
rotational stiffnesses; (2) the eccentricity e of the bearing
lines, shown in Figures 3(a) and 4; (3) the stiffness and the

stiffness ratios of subsequent piers, to identify whether the


bearing uplift effect is more severe in bridges with unequal
pier heights; (4) the stiffness of the foundation and (5) the
stiffness of the deck. The analysed bridge models were
considered to reflect potential alternative designs that a
Bridge Engineer may adopt.
Benchmark bridge (BB) model, shown in Figure 6,
took into account the shear and the rotational stiffness of
the bearings according to the Naeim and Kellys (1999)
bearing model. Detail 1 of Figure 6 shows the elastic K1,
the post-elastic K2 and the effective Keff shear stiffness of
the bearing model. The bearings of the BB are low
damping elastomeric bearings with a shear modulus
G 1 MPa. Damping was taken 5% of the critical one,
which corresponds to the expected damping for low
damping rubber bearings, while post-yield shear stiffness
was taken 50% of the elastic one, as shown in Table 2.
The effective stiffness was obtained for the two
different bearing geometries, i.e. B1 (at piers 1 3) 400
126(66) and B2 (at pier 4 and at abutments) 450 186
(110), considering that the target (i.e. design) shear strain
is 200% (Eurocode 8-2, EN 1998-2, 2005). The effective
stiffness of B1 and B2 was estimated to be 1713.60 and
1301.26 kN/m, respectively. The rotational stiffness of the
bearings, shown in Table 2 and in detail 1c of Figure 6,
was 998.4 and 1214.4 kN m/rad, respectively. The vertical
stiffness of the bearings under compression (Kc) was found
to be 303,470.37 kN/m for B1 and 291,660.51 kN/m for B2
(Naeim & Kelly, 1999).
Modelling of the bearings under tension was based on
the significant research of Yang et al. (2010). The
elastomeric bearings were considered to yield a tensile
stress equal to 2 MPa and a tensile strain equal to 3%. The
elastic tensile stiffness values (K t1 ) of bearings B1 and B2
are given in Table 2. The post-yield stiffness values K t2 ,
shown in Table 2, were taken equal to 6% of the initial
elastic ones (K t1 ) based on the v parameter described by
Yang et al. (2010). The bearings were set in two lines

Structure and Infrastructure Engineering

Vb

shear Vb

ub

ub
5

(a )

6
7
8

Nb
tension
continuity slab

vb
compr.

(a)

axial Nb
tension

2 rigid arm (eccentricity of cont. slab)


3 deck
4 rigid arm (bearing to deck's centroid)

(b)

5 bearing link

(a)shear, (b)axial, (c)rotational stiffness

(c) linear
rotational stiffness

6 rigid arm

(eccentricity e of bearings' support line)


7 pier cap
8 pier (frame elements)

(b)

P4
Detail 2

Detail 2

1 Pier

2 hinge of Pi
3 rigid arm

Detail 1

Pi (frame element)

Krxi, Kryi, Krzi (see Table 1)

Kxi,Kyi

P3

Kzi 4 spring elements


(foundation's flexibility)
2

z x
y
bending moment of the pier
M (kNm)

Downloaded by [82.24.31.176] at 06:48 11 December 2014

compr.

P1

P2

pier's bottom
cross section

25000
20000
15000

pier 4
pier 1
pier 3
pier 2

10000
5000
0
0

0.02
0.04
hinge rotation (rad)

0.06

Figure 6. The stick model of the BB. Detail 1, modelling of the pier deck connection; detail 2, modelling of the foundation and the
bending moment versus rotations, u, bilinear curves of potential piers hinges (left).

parallel to the transverse axis of the pier cap. Both bearing


lines 1 and 2 have a longitudinal distance e with respect to
the axis of the pier, as shown in Figure 6.
A bridge model, representing a typical design (TD),
was incorporated in the study. The TD model used a linear

model for the vertical stiffness of the bearing, i.e. stiffness


of the bearings under tension equal to the compressive one,
and small eccentricities e, as provided by AASHTO (2010)
figure 7.1-1 and Eurocode 8 Part 2 key figure of section
7.5. 4 (2005). The TD model used bearing shear and

8
Table 2.

S.A. Mitoulis
The stiffness values of the bearings of the BB (shown in detail 1 of Figure 5).

Downloaded by [82.24.31.176] at 06:48 11 December 2014

(a) Shear stiffness


(kN/m, nonlinear)

(b) Vertical stiffness


(kN/m, nonlinear)

(c) Bending stiffness


(kN m/rad, linear)

Bearing

Elastic
K1

Post-elastic
K2

Tensile elastic
K t1

Tensile post-elastic
K t2

Compressive
Kc

Kr

B1: 400 126(66)


B2: 450 186(110)

3129.8
2376.7

1564.9
1188.3

8377.6
10,602.9

502.7
636.2

303,470.4
291,660.5

998.4
1214.4

rotation stiffness values equal to the ones of the BB.


Modelling of isolated bridges based on the TD case
described above has been adopted for the analysis of
bridges by many researchers (Constantinou et al., 2011;
Dicleli, 2006; Hindi and Dicleli, 2006; Kappos, Saiidi,
Aydinoglu, & Isakovic, 2012; Mwafy, Elnashai, & Yen,
2007).
All the analyses used 3D stick models, which followed
the guidelines of Kappos et al. (2012, Chapter 12).
Figure 6 shows the stick model of the BB. The deck was
modelled by frame elements. The piers were modelled by
frame elements connected in series to nonlinear rotational
springs that modelled the potential plastic hinges, as
shown in detail 2 of Figure 6. The moment rotation (M
u) curves of the piers, which are illustrated in Figure 6,
were calculated by means of the RCCOLA-90 software
(Kappos, 2002). The post-elastic stiffness of the piers was
assumed to be 2% of its initial elastic one.
The flexibility of the foundation was taken into
account by assigning linear and rotational springs in the
two horizontal and in the vertical direction of the
foundation. The soil spring values, given in Table 3,
were directly adopted by the geotechnical in situ tests of
the as-built bridge. The expressions for the subgrade
reaction modulus for the soil were obtained from Poulos
and Davis (1980). The bridge models were subjected to
longitudinal seismic excitations. Artificial accelerograms
that were compatible to ground types B (vs,30 < 580 m/s,
N SPT . 50 and c u . 250 kPa) and C-dependent
(vs,30 < 270 m/s, NSPT < 32 and cu < 160 kPa) Eurocode
8-1 (EN 1998-1, 2005) elastic spectra were used (Sextos,
Kappos, & Pitilakis, 2003), while the accelerograms were
scaled to three levels of seismic excitation representing
peak ground accelerations (PGAs) of 0.25, 0.5 and 0.75 g.
Table 3.

The nonlinear response of bridge models was analysed


using the FEM code SAP 2000 ver. 14.2.0 (Computers and
Structures Inc. [CSI], 2010). NLDTHA was implemented
and the average acceleration method Newmark was
chosen (Chopra, 1995). The mass and stiffness proportional damping was chosen and critical damping ratios
equal to 5% and 4% were considered for the first and the
second longitudinal periods of the analysed bridge
systems. The BB was found to have a fundamental
longitudinal period of 2.6 s, while the modal participation
factor was found to be 0.81.
4. Results
4.1. Identification of bearings uplift effect
The development of the bearings uplift effect was
confirmed by calculating the time histories of the vertical
displacements of the bearings and the vertical displacements of the deck upwards, when the BB was subjected to
a longitudinal seismic excitation. Figure 7(a) shows the
time histories of the longitudinal deck movements, when
the bridge was subjected to a ground motion with a PGA
equal to 0.50 g, i.e. almost three times larger than the
actual design earthquake. The artificial accelerogram used
in this case corresponded to Eurocode 8-1 (EN 1998-1,
2005) ground type B. The figure illustrates that the
maximum horizontal displacement of the deck along x
axis is almost 457 mm, while the one along x direction is
408 mm. These displacements were found to induce
rotations to the pier cap of pier 4 equal to w1 25
1023 rad (clockwise as shown in Figure 7(d)) and
w2 22 1023 rad (counter-clockwise as illustrated in
Figure 7(e)). The corresponding rotations of the deck
about its transverse axis were found to be 4 1023 rad

The stiffness values of the foundation.

Kx (kN/m)
Ky (kN/m)
Kz (kN/m)
Krx (kN m/rad)
Kry (kN m/rad)
Krz (kN m/rad)

Pier 1

Pier 2

Pier 3

Pier 4

1.21 106
1.09 106
3.14 106
38.08 106
36.56 106
0.59 106

0.69 106
0.75 106
3.06 106
34.90 106
31.88 106
0.59 106

12.84 106
15.83 106
4.40 106
66.26 106
63.31 106
0.61 106

16.08 106
16.84 106
4.40 106
66.68 106
66.02 106
0.61 106

Structure and Infrastructure Engineering

long. displac. (mm)


ud

(a)

600
200
200
400

vert. displac. (mm)


vd

ud2max negative displ. (x)

408mm
5

10
time (s)

15

left
bearing
(compr.)

u p1

right
bearing
(tension)

20

40
due to pier
cap rotations

30

vd1

due to vert. seismic


component

vd2

20
(e)

10

ud2max negative displ. (-x)


d2

d2

0
uz,EZ
0

vert. displac. vb(mm)


compr. tension

d1

10

Downloaded by [82.24.31.176] at 06:48 11 December 2014

d1

time (s)

(c)

(d)
u d1 max positive displ. (+x)

ud1max positive displ. (+x)

457mm

400

600
(b)

80

5
61mmright bearing

60

10

15

20

left
bearing
(tension)

left bearing

up2

right
bearing
(compr.)

54mm

40
20
0
20

10
time (s)

15

20

Figure 7. Time histories of (a) the longitudinal deck movements; (b) the vertical deck movements due to the rotations of pier 4 cap beam
(dashed line) and due to the vertical component of the seismic action (continuous line); (c) the vertical displacements of the bearings on
the left (line 1) and on the right (line 2); (d) the deformed pier deck connection when the deck displaces on the right and (e) on the left
(PGA 0.5 g and ground type B).

(clockwise) and 3.8 1023 rad (counter-clockwise).


Hence, the rotations of the deck are one-sixth of those of
the pier cap.
Figure 7(b) illustrates the time history of the vertical
deck displacements caused by the rotations of pier 4 cap
beam (dashed line). The maximum vertical displacements
of the deck were calculated to be almost vd1 30 mm and
vd2 27 mm. Displacements vd1 and vd2 were found to
occur together with the corresponding maximum positive
( x) and maximum negative ( x) horizontal displacements of the deck, shown in Figure 7(a), and the w1 and w2
rotations of the pier caps shown in Figure 7(d),(e),
respectively. For comparison, the deck movements were
also calculated for the vertical seismic excitation
(continuous line). It is observed that the deck is dislodged
upwards due to the rotations of the pier cap. Vertical deck
displacements due to the rotations of the pier cap were
found to be larger than the ones induced by the vertical
component of the seismic excitation, in most cases. Also,
based on analysis of the vertical deck motion by fast
Fourier transform, the displacements vd, which are caused
by the pier cap rotations, were found to exhibit lower

frequencies than the ones of the vertical seismic


component, as shown in Figure 7(b).
The pier cap clockwise rotation w1 and the deck
movement vd1 upwards induce a tensile displacement to
the bearings at line 2 equal to 61 mm (bearings on the right
side of the pier cap). This displacement is shown in
Figure 7(c),(d). Figure 8 shows in detail the deformed
shape of the pier cap deck connection, when the deck
exhibits the displacement ud1 , which corresponds to a
bearing shear strain 1s 1.28. At the same time, the
bearings on the left (line 1) are compressed, as shown in
Figure 7(c) with the continuous line and in Figure 8. The
opposite was found to be valid for the counter-clockwise
rotation w2 of the pier cap. When the deck exhibits its
maximum negative displacement ud2 along x axis, the
bearings at line 1 were found to be uplifted by 54 mm, as
shown in Figure 7(c),(e).
The offset compressive displacement of the bearings
due to quasi-permanent vertical loads, shown in Figure 2,
was removed from all the results illustrated in the study,
such as the initial vertical displacement of the bearings,
which is negative (downwards), was set equal to zero, i.e.

10

S.A. Mitoulis
3

410 rad
31mm
36.5mm
before
and after
seismic excitation

4 10 rad
457mm

451mm
line 1

37mm

310mm

line 2
3
310mm 25 10 rad

before
and after
seismic excitation

310mm
37mm
3
2510 rad

2510 rad
e e=1.5m

support line 1:
total shear displ.: 141mm (e =1.28)
total vert. displ. : 0.5mm (compressive)
3
total rotation : 2110 rad (clockwise)

410 rad
24mm
451mm

support line 2:
total shear displ.: 141mm (es =1.28)
total vert. displ. : 61mm (tensile)
3
total rotation : 2110 rad (clockwise)

the bearing vertical displacements were measured from the


deck dead load condition. Maximum tensile displacements
of bearings at line 1 (61 mm) and line 2 (54 mm)
correspond to axial tensile strains 56% and 49% that are
greater than both the tensile strains at the onset of
cavitation and at yield, which is of the order of 3%.
4.2. Influence of bearing uplift on the response of
bridges
The response of the BB model was compared with one of
the TD case, which did not account for the eccentricities e
of the bearings nor for the nonlinear vertical response of
the bearings. The comparison was carried out on the basis
of shear and tensile displacements of the bearings and
bending moments of the pier. Figure 9(a) depicts the
comparison of the maximum shear displacements of the
bearings, while Figure 9(b) illustrates the maximum
tensile deformations of the isolators for the two different
modelling alternatives. A seismic input motion with a
PGA of 0.50 g was analysed that was compatible to
Eurocode ground types B and C.
The figure shows that the shearing of bearings is not
strongly influenced by the model used for the vertical

Ab.1

response of the bearings. The maximum difference


between the shear displacements of the bearings was
observed for the bearings at abutments and was almost
14%. Conversely, the tensile displacements of the isolators
differ significantly between the two different bridge
models. Figure 9(b) shows that the bearings of the bridge
that represent a TD responded with vertical tensile
displacements from 13 mm (at abutments) up to 24 mm at
piers 1 and 4. The bearings of the BB yielded tensile
movements from 12 mm (at abutments) up to 58 and
61 mm at piers 1 and 4, respectively.
Hence, the tensile displacements of the bearings at
shorter piers are increased by a factor of almost 2.4 when
nonlinear modelling of the bearings is adopted. A further
investigation showed that bearings on shorter piers are
more prone to exhibit tensile stresses, than the ones on tall
piers, due to the fact that the rotation of shorter piers
foundation induces a clear and almost equal rotation to the
pier cap that, in turn, induces tensile stresses to bearings.
Conversely, the flexural deflections of taller piers and the
resulting rotation of the pier cap were found to be
effectively restrained by the vertical stiffness of the
bearings, while the rotation of their foundations was much
smaller.
(b)

benchmark bridge (BB)


ground type B
ground type C

tensile displacements (mm)

(a) 1000
900
800
700
600 14%
500
400
300
200
100
0

shear displacements (mm)

Downloaded by [82.24.31.176] at 06:48 11 December 2014

Figure 8. The deformed shape of the pier cap deck connection (pier 4) when the deck displaces with its maximum positive (x)
displacement.

typical design case (TD)


ground type B
ground type C

bearings of:
Pier1 Pier2 Pier3 Pier4

Ab.2

100
90
80
70
60
50
40
30
20
10
0

BB

ground type B
ground type C

TD ground type B
ground type C

Ab.1

bearings of:
Pier1 Pier2 Pier3 Pier4

Ab.2

Figure 9. (a) The maximum shear displacements of the bearings of the benchmark and the TD case and (b) the maximum tensile
displacements of the bearings (PGA 0.5 g).

Analyses of bridges for lower (0.25 g) and higher


(0.75 g) seismic excitations yielded similar trends. For a
PGA of 0.25 g, the bearings at pier 4 responded with
maximum tensile movements 30 or 11 mm, when the
bridge model took into account (BB) or neglected (TD
case) the vertical nonlinear response of the bearings,
respectively. The corresponding displacements were 86
and 36 mm for a PGA of 0.75 g. The soil type did not seem
to influence strongly the shear or the tensile displacements
of the BB and the TD elastomeric bearings, as shown in
Figure 9(a),(b). Hence, vertical tensile displacements of
the bearings are significantly underestimated when a linear
model is used for modelling the vertical response of the
isolators and when the eccentricities of the bearings, with
respect to the axis of the piers, are neglected. Hence, in
most cases, the bearings seem to exceed, by far, the
yielding tensile strain of 3%.
The influence of the bearing uplift mechanism on the
seismic response of the piers was compared for the two
different bridge models employed in the study. Figure 10
(a) illustrates the maximum bending moments at the
bottom section of the piers, shown in detail 2 of Figure 6,
when the bridge models were subjected to a ground motion
with a PGA of 0.50 g for ground type B. The comparison
reveals that the maximum bending moments at the bottom
section of the piers of the BB are up to 13% smaller when
the vertical model of the bearings is taken into account, as
shown in Figure 10(a). The opposite was found to be valid
for the bending moments at the top of the piers, shown in
Figure 10(b).
The piers of the bridge representing the TD case,
which did not take into account the eccentricity and the
vertical nonlinear response of the bearings, exhibited
relatively small bending moments at the top (i.e. 1136.1,
768.6, 902.9 and 1438.6 kN m at piers 1 4, respectively).
The corresponding bending moments of the piers of the
BB were 6197.7, 4043.6, 4306.5 and 5805.3 kN m. The
results for the different levels of seismic excitation 0.25 g

20000
15000
10000
5000
0
P2

P3

Pier bottom section

4.3. Bearing uplift effect for different bridge


configurations
Specific design parameters of bridges ERS were
parametrically analysed to assess their influence on the
seismic displacements of bearings in the vertical direction.
The parametric study utilised the geometry and the
modelling of the BB, i.e. nonlinear modelling of the
bearings was incorporated for all the analyses. The
influence of the eccentricity e on the magnitude of bearing
uplift displacements was studied. The same bridge model
was reanalysed for different eccentricities. Common
values of the eccentricities were calculated based on (1)
the distance of the precast I-beam faces that ranged in
practice from 200 to 800 mm; (2) the distance between the
end of the beam and the edge of the bearing that ranged
from 100 to 400 mm and (3) the dimensions of the bearings
in precast I-beam decks that ranged from 400 to 800 mm.
Figure 11(a) illustrates the tensile deformations of the
bearings at abutments and piers for eccentricities 0.5, 1.0
and 1.5 m. The displacements were calculated for ground
acceleration 0.25 g and for ground type B. The figure
shows that the larger the eccentricity of the bearings with
respect to the axis of the pier, the larger are the vertical
tensile displacements of the bearings. This was attributed
to the fact that, for the same pier cap rotation about its
transverse axis, the displacement of the pier cap vp < we,
which is shown in the detail of Figure 3(a), increases when
the eccentricity e is increased. Hence, the tensile
displacements of the bearings, which are increased when
vp increases, tend to be larger at larger eccentricities.

(b) 7000

typical design case (TD)


benchmark bridge (BB)

P1

11

and 0.75 g were similar, namely the bending moments at


the bottom of the piers were higher at the TD, while the
model of the BB yielded larger bending moments at the top
of the piers. Based on these results, it seems that the top
sections of the piers should be checked and reinforced
properly against bending moments.

bending moment (kNm)

(a) 25000

bending moment (kNm)

Downloaded by [82.24.31.176] at 06:48 11 December 2014

Structure and Infrastructure Engineering

P4

6000

typical design case (TD)


benchmark bridge (BB)

5000
4000
3000
2000
1000
0
P1

P2
P3
Pier top section

P4

Figure 10. The maximum bending moments of the piers of the BB and the TD case: (a) bottom section and (b) top section (PGA 0.5 g,
soil type B).

12

S.A. Mitoulis
(b) 100
90
80
70
60
50
40
30
20
10
0
tensile displacements (mm)

tensile displacements (mm)

(a) 100
90
80
70
60
50
40
30
20
10
0

1.5m
e= 1.0m
0.5m

1.5m
e= 1.0m
0.5m

bearings of:

bearings of:

Ab.1 Pier1 Pier2 Pier3 Pier4 Ab.2

Ab.1 Pier1 Pier2 Pier3 Pier4 Ab.2

(a) 100
90
80
70
60
50
40
30
20
10
0

0.2Kbo Kbo 5Kbo

displacements of bearings that are eccentrically placed


with respect to the axis of the pier cap tend to be larger
when the seismic excitation is increased and when the
eccentricity e is larger. Observable was that the yielding
axial strain of bearings under tension (3%) was exceeded
in all bearings.
The influence of different bearing dimensions,
corresponding to variable shear, axial and bending
stiffness values, on the uplift displacements of the
isolators was studied by analysing alternative models for
the BB having different elastomeric bearing properties.
Realistic design of the bearings was carried out based on
the study by Manos, Mitoulis, and Sextos (2012). Figure 12
(a) shows the uplift (tensile) displacements of the bearings
calculated for the bridge models that utilised three
different elastomeric bearing stiffnesses. The stiffness of
the bearings ranged from 0.2Kbo to 5Kbo, where Kbo
represents the stiffness of the bearings of the BB model.
Figure 12(a) illustrates the tensile deformation of the
bearings for the normal level of the seismic excitation (i.e.
0.25 g) and for ground type B. The maximum tensile
displacements of the bearings at pier 2 were found to be

(b) 100
90
80
70
60
50
40
30
20
10
0
tensile displacemetns (mm)

Again, the largest uplift displacements of the bearings


were concentrated above the shorter piers 1 and 4, which
was also observed in Figure 9(b).
Bearings at pier 4, which was the shortest pier among
the others, were found to respond with a maximum tensile
displacement equal to 14, 25 and 31 mm when the
eccentricities were 0.5, 1.0 and 1.5 m, respectively. Hence,
the increase in the eccentricity by a factor of 2 (from 0.5 to
1.0 m) led to an almost 78% increase in the tensile
displacement of the bearings. Further increase in the
eccentricity by a factor of 3 (from 0.5 to 1.5 m) increased
the tensile displacement of the bearings by a factor of 2.2
(from 14 to 31 mm). The influence of the eccentricity on
the tensile displacements of the bearings was found to
have similar trends when the high seismic excitation of
0.75 g was considered, as shown in Figure 11(b). The
vertical tensile displacements of the bearings at pier 4 were
found to be 28, 76 and 96 mm when the eccentricities were
0.5, 1.0 and 1.5 m, respectively. The last increases in the
eccentricities by factors of 2 (from 0.5 to 1.0 m) and 3
(from 0.5 to 1.5 m) yielded increases in the displacements
by factors of 2.7 and 3.4, respectively. Hence, the uplift

tensile displacemetns (mm)

Downloaded by [82.24.31.176] at 06:48 11 December 2014

Figure 11. The influence of the eccentricity e on maximum tensile displacements of the bearings of the BB: (a) normal seismic
excitation 0.25 g and (b) high seismic excitation 0.75 g (ground type B).

0.2Kbo Kbo 5Kbo

bearings of:

bearings of:

Ab.1 Pier1 Pier2 Pier3 Pier4 Ab.2

Ab.1 Pier1 Pier2 Pier3 Pier4 Ab.2

Figure 12. The influence of the bearings stiffness on their maximum tensile displacements. The bridge models were subjected to (a)
normal seismic excitation 0.25 g and (b) high seismic excitation 0.75 g (ground type B).

Structure and Infrastructure Engineering

Case 1 Case 2 Case 3

Case 4

(b) 100
90
80
70
60
50
40
30
20
10
0

Case 1 Case 2

bearings at:
Ab.1 Pier1 Pier2 Pier3 Pier4 Ab.2

Case 3

Case 4

tensile displacements (mm)

tensile displacements (mm)

(a) 100
90
80
70
60
50
40
30
20
10
0

13

bearings at:
Ab.1

Pier1 Pier2 Pier3 Pier4

Ab.2

Downloaded by [82.24.31.176] at 06:48 11 December 2014

Figure 13. The maximum tensile displacements of the bearings for bridge models with either equal pier heights (Case 1 all piers have
height 10 m) or variable piers heights. Case 2 has piers with heights 10, 20, 20 and 10 m; Case 3 has 10, 20, 15.7 and 12.5 m and Case 4 has
10, 17.4, 25 and 12 m: (a) PGA 0.5 g and (b) PGA 0.75 g.

21, 19 and 5 mm when bearings with stiffnesses of 0.2Kbo,


Kbo and 5Kbo were considered, respectively. Uplift
displacements equal to 18, 31 and 12 mm were calculated
for the bearings at pier 4 when bearings with stiffnesses of
0.2Kbo, Kbo and 5Kbo were used, respectively.
Figure 12(b), which represents the uplift mechanism at
larger seismic excitations, i.e. 0.75 g, shows that the tensile
displacements of the isolators at pier 2 were 63, 55 and
15 mm when bearings with stiffnesses of 0.2Kbo, Kbo and
5Kbo were used, respectively. The corresponding displacements of bearings at pier 4 were 55, 96 and 37 mm.
Hence, it seems that stiffer bearings reduce, in most cases,
the tensile displacements of the isolators. This was found
to be valid for the bearings at piers of great heights, i.e.
piers 2 and 3. However, the uplift displacements of
bearings at shorter piers 1 and 4 were either increased,
when the stiffness of the bearings was increased from
0.2Kbo to Kbo, or reduced when bearing stiffness was
increased from Kbo to 5Kbo.
Having observed in Figures 9, 11 and 12 that the uplift
mechanism of the bearings is affected significantly by the
heights of the piers, bridge models with different heights
of the piers were analysed. Bridges having pier with equal
heights and bridge systems with significant variations in
the pier heights were studied. Stiffness ratios were then
calculated for all the piers with respect to the stiffness of
the most flexible pier of the bridge model. The following
bridge systems were analysed, while keeping the pier cross
sections identical to the ones of the BB: (Case 1) all pier
heights equal to 10 m; (Case 2) piers 1 4 with heights 10,
20, 20 and 10 m, respectively, corresponding to stiffness
ratios of 8, 1, 1 and 8; (Case 3) pier heights 10, 20, 15.7
and 12.5 m corresponding to stiffness ratios of 8, 1, 2 and 4
and (Case 4) pier heights 10, 17.4, 25 and 12 m
corresponding to stiffness ratios of 16, 3, 1 and 9.
Figure 13(a) illustrates the tensile displacements of the
bearings at the abutments and the piers for bridge Cases

1 4 described above, for a PGA of 0.5 g. The figure shows


that the bearing uplift is more intense in bridges whose
piers have equal heights, i.e. in bridges with evenly
distributed seismic actions under a longitudinal earthquake
motion. The reason why bridge model of Case 1 resulted in
the maximum tensile displacements of the bearings was
that all the pier caps exhibited their maximum rotation
about a transverse axis (almost equal to 16 103 rad)
when the deck moved longitudinally with its maximum
seismic displacement of 240 mm. The deck in this case
was found to be uplifted with an almost uniform vertical
displacement upwards vd 19 mm, due to the inability of
the bearings under compression to absorb the vertical
movement of the piers upwards vp 22 mm, shown in the
detail of Figure 3(a).
Accordingly, bearings under tension were pulled off by
a maximum tensile displacement of 55 mm. The bridge
model of Case 4, which corresponds to the bridge with the
largest discrepancy in pier heights and stiffness ratios,
yielded the smallest bearings tensile displacements at pier
3 among the other bridge cases. This was found to occur
due to the fact that under the maximum longitudinal
displacement of the deck that was equal to 301 mm, the
pier cap of the highest pier 3 exhibited a rotation of 12
103 rad about its transverse axis, while the rotation of
shorter pier 1 was found to be almost 20 103 rad. These
values correspond to the total rotation of the piers, i.e. they
include both the rotation of the foundation and the flexural
deflection of the pier.
Consequently, bearings at pier 3 experienced tensile
displacements almost 35 mm that are smaller than the 55mm tensile deformation of bearings at pier 1. Hence, the
uplift mechanism of the bearings is more critical in bridges
with equal pier heights, at least for the cases studied. The
trends of bearing uplift displacements were found to be
similar when the high seismic action, i.e. 0.75 g, was
considered. Figure 13(b) shows that, indeed, bridge

Downloaded by [82.24.31.176] at 06:48 11 December 2014

14

S.A. Mitoulis

bearings of case 1 exhibited the maximum tensile


displacements.
The uplift mechanism of the bearings was further
studied with respect to the cross sections of the piers, the
foundations and the deck. The BB was reanalysed for more
flexible and for stiffer piers, having a stiffness of 0.2 and
five times the stiffness Kpo of the piers of the BB. The
alternative stiffness values were considered to represent
realistic pier design alternatives of isolated bridges, which
correspond to multi-column piers (flexible piers with a
stiffness of 0.2Kpo) and hollow rectangular piers (stiff
piers with a stiffness of 5Kpo). The analyses showed that
the stiffer the piers, i.e. the stiffer the overall bridge
system, the smaller the bearing uplift displacements. This
was found to occur due to the fact that the maximum
displacement of the deck and hence the resulting rotations
of the pier caps become smaller when the overall bridge
system is stiffer. Thus, the vertical displacements of the
pier cap vp are effectively reduced.
Suggestively, the increase in the piers stiffness from
Kpo to 5Kpo yielded about half uplift displacements of the
isolators. However, the above outcome that relates the total
stiffness of the bridge with the vertical displacements of
the bearings should not be related to the outcome, which is
relevant to bridges with short and tall, i.e. stiffer and more
flexible piers. The last outcome related the interaction
between the piers of different stiffnesses in the same
bridge and not the development of bearing vertical
displacements for different total bridge stiffnesses.
Similarly, stiffness values of different foundations
were parametrically analysed considering a stiffness of 0.2
and five times stiffer than the foundations of the BB (Kfo).
Results showed that the stiffer the foundations, the smaller
the tensile displacements of the bearings become.
However, the stiffness of the foundations did not yield a
remarkable influence on the bearing uplift displacements.
It was found that the variations in the tensile displacements
of the bearings was marginal of the order of 10% for
the three different levels of foundation stiffness, i.e.
0.2Kfo, Kfo and 5Kfo. Relatively low was also the
influence of the deck stiffness on the tensile displacements
of the bearings. A stiffer deck, namely a deck that has a
moment of inertial about its transverse axis greater than
the deck of the BB, was found to amplify the uplift
mechanism of the bearings. Indicatively, the analyses
showed that a box-girder superstructure, being almost
three times stiffer than the deck of the BB, yielded almost
26% increase in the uplift displacements of the bearings.
5.

Conclusions

An uplift mechanism of anchored elastomeric bearings


was described. The equilibrium of the seismic actions
during bearings unseating was derived for an isolated
bridge subjected to seismic excitation in the longitudinal

direction. Typical bearing placement was adopted


considering two lines of support parallel in the transverse
direction (y) of the pier cap, while having an eccentricity e
with respect to the axis of the pier. The analysis showed
that the isolators experience compressive and significant
tensile (uplift) displacements due to the rotations of the
pier caps about its transverse axis. The magnitude of the
bearing uplift displacements was studied by taking into
account the nonlinear response of the isolators in the
vertical direction. The study has led to the following
conclusions.
The maximum tensile displacements of anchored
bearings were found to occur when the deck displaced with
its maximum longitudinal (along x) displacement during
an earthquake. The analysis of the BB, which was
subjected to a seismic motion with a PGA of 0.5 g, showed
that the maximum positive (457 mm, along x) and
negative (408 mm along x) displacements of the deck
induced the maximum clockwise and counter-clockwise
rotations to the pier caps, respectively. These rotations
forced the bearings to receive tensile displacements 61 mm
(bearing at the right side of the pier cap) and 54 mm
(bearing at the left side of the pier cap). Bearings on
shorter piers were found to exhibit larger tensile movements than the bearings on taller piers.
Larger eccentricities e of the bearings, with respect to
the axis of the pier, were found to increase the tensile
displacements of the bearings. Bearings on the shortest
pier 4 were found to respond with maximum tensile
displacements equal to 14, 25 and 31 mm when the
eccentricities were 0.5, 1.0 and 1.5 m, respectively. These
displacements were obtained for a PGA equal to 0.25 g and
for ground type B. Hence, the increase in the eccentricity
by a factor of 2 or 3 increased the uplift displacements by
1.8 and 2.2 times. The corresponding factors were 2.7 and
3.4 when the higher seismic action was adopted, i.e. 0.75 g.
Analyses of the BB were carried out for three different
bearing types having axial stiffnesses 0.2Kbo, Kbo and
5Kbo with Kbo being the stiffness of the isolators of the
BB. The analysis for a PGA of 0.25 g showed that the
maximum tensile displacements of the bearings at pier 2,
which is the tallest pier of the bridge, were found to be 21,
19 and 5 mm for the three different bearing stiffnesses.
Moreover, bridge models with equal pier heights and
bridges with uneven distribution of seismic actions were
analysed. Analyses showed that the uplift mechanism is
more critical when the piers have the same heights. The
last was found to be attributed to the fact that the
maximum longitudinal displacement of the deck induces
almost the same and synchronous rotations of the pier
caps, which in turn induce tensile strains to the bearings.
Bridge models with different cross sections of the
piers, the foundations and the deck were analysed.
Analyses showed that the bridge model that utilised
hollow rectangular piers, having stiffness almost five times

Downloaded by [82.24.31.176] at 06:48 11 December 2014

Structure and Infrastructure Engineering


greater than the circular piers of the BB, decreased the
maximum bearing uplift displacements by a factor of 2.1.
Accordingly, the increase in the stiffness of the foundation
by a factor of 5 led to a reduction in the bearing uplift
displacements by almost 10%. Conversely, the increase in
the deck stiffness having a moment of inertia three times as
that of the BB, led to increased uplift displacements of the
bearings by almost 26%.
The study showed that the nonlinear response of bridge
bearings in the vertical direction should be modelled when
the bearings are placed eccentrically with respect to the
axis of the pier. Uplift effects and consequent cavitation of
the bearings were found to be expected in all bridge cases
studied in this paper. The severity of the effect could be
identified by checking the bearings of the shorter piers.
A case-dependent analysis is required for the design of the
bridge isolation system taking into account the different
geometries of the deck and the substructure, as well as the
combination of the three components of the seismic action.
Disclosure statement
No potential conflict of interest was reported by the author(s).

Note
1.

Prior to the final parametric study, different stiffnesses of the


continuity slab were tested analytically to define whether the
stiffness of the connecting slab allows for different rotations
of the bearings. The results of the analyses showed that the
discrepancy of the bearings rotations due to the flexibility of
the deck slab is negligible.

References
Akiyama, M., Frangopol, D.M., & Mizuno, K. (2014).
Performance analysis of Tohoku-Shinkansen viaducts
affected by the 2011 Great East Japan earthquake. Structure
and Infrastructure Engineering, 10, 1228 1247.
American Association of State Highway and Transportation
Officials [AASHTO] (2010). Guide specifications for seismic
isolation design (3rd ed.). Washington, DC: Author.
American Association of State Highway and Transportation
Officials [AASHTO] (2013). LRFD bridge design specifications. Customary U.S. units, 6th ed., with 2013 interim
revisions. Washington, DC: Author.
Aria, M., & Akbari, R. (2013). Inspection, condition evaluation
and replacement of elastomeric bearings in road bridges.
Structure and Infrastructure Engineering, 9, 918 934.
Artieda, C.C.M., & Whittaker, A.S. (2010). Theoretical studies
of the XY-FP seismic isolation bearing for bridges. Journal
of Bridge Engineering, 15, 631 638.
Buckle, I., Yen, W.-H.(.P.), Marsh, L., & Monzon, E. (2012).
Implications of bridge performance during Great East Japan
Earthquake for U.S. seismic design practice. In Proceedings
of the International Symposium on Engineering Lessons
Learned from the 2011 Great East Japan Earthquake, March
1 4, 2012, Tokyo (pp. 1363 1374). Tokyo: Japan
Association for Earthquake Engineering. Retrieved from
http://nisee.berkeley.edu/elibrary/getpkg?id201203281

15

BS EN 15129 (2009). Anti-seismic devices. British Standards


International. Retrieved from http://shop.bsigroup.com/ProductDetail/?pid=000000000030129844
California Department of Transportation [CalTrans] (1999).
Bridge memo to designers (20-1) Seismic design
methodology. Sacramento, CA: Author.
Chang, K.-C., Chang, D.-W., Tsai, M.-H., & Sung, Y.-C. (2000).
Seismic performance of highway bridges. Earthquake
Engineering and Engineering Seismology, 2, 55 77.
Chu-Chieh, L., Hung, J.H.-H., Liu, K.-Y., & Chai, J.-F. (2010).
Reconnaissance observation on bridge damage caused by the
2008 Wenchuan (China) earthquake Chu-Chieh. Earthquake
Spectra, 26, 1057 1083.
Chopra, A.K. (1995). Dynamics of structures Theory and
applications to earthquake engineering. Englewood Cliffs,
NJ: Prentice-Hall International Series in Civil Engineering
and Engineering Mechanics.
Computers and Structures Inc. (CSI) (2010). SAP 2000
Nonlinear version 14.2.0, users reference manual. Berkeley,
CA: Author.
Constantinou, M.C., Kalpakidis, I., Filiatrault, A., & Ecker Lay,
R.A. (2011). LRFD-based analysis and design procedures
for bridge bearings and seismic isolators (MCEER-110004). Buffalo, NY: Multidisciplinary Center for Earthquake
Engineering Research. Retrieved from http://mceer.buffalo.
edu/pdf/report/11-0004.pdf
Dicleli, M. (2006). Performance of seismic-isolated bridges in
relation to near-fault ground-motion and isolator characteristics. Earthquake Spectra, 22, 887 907.
Dorfmann, A., & Burtscher, S.L. (2000). Aspects of cavitation
damage in seismic bearings. Journal of Structural Engineering, 126, 573 579.
Earthquake Engineering Research Institute [EERI] (1995a).
Northridge earthquake reconnaissance report. Oakland, CA:
Author.
Earthquake Engineering Research Institute [EERI] (1995b). The
Hyogo-Ken Nanbu earthquake reconnaissance report. Oakland, CA: Author.
Earthquake Engineering Research Institute [EERI] (2001). 1999
Chi-Chi, Taiwan, earthquake reconnaissance report. Oakland, CA: Author.
Earthquake Engineering Research Institute [EERI] (2011).
Learning from earthquakes bridge performance in the Mw
9.0 Tohoku, Japan, earthquake of March 11, 2011. Oakland,
CA: Author.
EN 1337-3 (2005). Structural bearings Part 3: Elastomeric
bearings. Brussels: European Committee for Standardization.
EN 1992-1-1 (2004). Eurocode 2: Design of concrete structures,
Part 1: General rules and rules for buildings. Brussels:
European Committee for Standardization.
EN 1998-1 (2005). Eurocode 8: Design of structures for
earthquake resistance, Part 1: General rules, seismic actions
and rules for buildings. Brussels: European Committee for
Standardization.
EN 1998-2 (2005). Eurocode 8: Design of structures for
earthquake resistance, Part 2: Bridges. Brussels: European
Committee for Standardization.
Hamzeh, O., Tassoulas, J., & Becker, E. (1998). Behavior of
elastomeric bridge bearings: Computational results. Journal
of Bridge Engineering, 3, 140 146.
Hindi, R., & Dicleli, M. (2006). Effect of modifying bearing fixities
on the seismic response of short- to medium-length bridges
with heavy substructures. Earthquake Spectra, 22, 6584.
Imbsen, R.A. (2007). Proposed AASHTO guide specifications for
LRFD Seismic Bridge Design-Subcommittee for seismic

Downloaded by [82.24.31.176] at 06:48 11 December 2014

16

S.A. Mitoulis

effects on bridges, T-3. Imbsen Consulting. Retrieved from


http://bridges.transportation.org/Documents/2007BallotSeismicGuidelines.pdf
Iwabe, N., Takayama, M., Kani, N., & Wada, A. (2000).
Experimental study on the effect of tension for rubber bearings.
In Proceedings of the 12th WCEE-World Conference on
Earthquake Engineering, Auckland, New Zealand, paper 1290.
International Association for Earthquake Engineering.
Retrieved from http://www.iitk.ac.in/nicee/wcee/article/1290.
pdf
Japan Road Association [JRA] (1997). Manual for seismic design
of highway bridges. Tokyo: Author.
Japan Road Association [JRA] (2002). Chapter 1: Seismic design
specifications for highway bridges. Tokyo: International
Institute of Seismology and Earthquake Engineering, Japan
Road Association.
Kappos, A.J. (2002). RCCOLA-90: A microcomputer program
for the analysis of the inelastic response of reinforced
concrete sections (Report). Thessaloniki: Department of
Civil Engineering, Aristotle University of Thessaloniki.
Kappos, A.J., Saiidi, M.S., Aydinoglu, M.N., & Isakovic, T.
(2012). Seismic design and assessment of bridges, inelastic
methods of analysis and case studies. Geotechnical,
Geological and Earthquake Engineering, Vol. 21. The
Netherlands: Springer.
Katsaras, C.P., Panagiotakos, T.B., & Kolias, B. (2009). Effect of
torsional stiffness of prestressed concrete box girders and
uplift of abutment bearings on seismic performance of
bridges. Bulletin of Earthquake Engineering, 7, 363 375.
Kawashima, K. (2012). Damage of bridges due to the 2011 Great
East Japan earthquake. In Proceedings of the International
Symposium on Engineering Lessons Learned from the 2011
Great East Japan Earthquake, March 14, 2012, Tokyo, Japan
(pp. 82101). Tokyo: Japan Association for Earthquake
Engineering (JAEE). Retrieved from http://www.jaee.gr.jp/
event/seminar2012/eqsympo/pdf/papers/140.pdf
Kelly, J.M. (2003). Tension buckling in multilayer elastomeric
bearings. Journal of Engineering Mechanics, 129,
1363 1368.
Kelly, J.M., & Takhirov, S.M. (2007). Tension buckling in
multilayer elastomeric isolation bearings. Journal of
Mechanics of Materials and Structures, 2, 1591 1605.
Kitahara, T., Kajita, Y., & Kitane, Y. (2012). Investigation on the
damage cause of the bridge rubber bearings in the 2011 off
the Pacific coast of Tohoku earthquake. In Proceedings of the
15th WCEE World Conference on Earthquake Engineering, Lisbon, Portugal, paper No. 2815. International
Association for Earthquake Engineering. Retrieved from
http://www.iitk.ac.in/nicee/wcee/article/WCEE2012_2815.
pdf

Kunde, M.C., & Jangid, R.S. (2003). Seismic behavior of isolated


bridges: A state-of-the-art review. Electronic Journal of
Structural Engineering, 3, 140 170.
Manos, G.C., Mitoulis, S.A., & Sextos, A. (2012). A knowledgebased software for the preliminary design of seismically isolated
bridges. Bulletin of Earthquake Engineering, 10, 10291047.
McDonald, J., Heymsfield, E., & Avent, R.R. (2000). Slippage of
neoprene bridge bearings. Journal of Bridge Engineering, 5,
216 223.
Mitoulis, S.A. (2012). Seismic design of bridges with the
participation of seat-type abutments. Engineering Structures,
44, 222 233.
Mitoulis, S.A., Muhr, A., & Ahmadi, H. (2014). Uplift of
elastomeric bearings in isolated bridges A possible
mechanism: Effects and remediation. In Proceedings of the
2nd European Conference on Earthquake Engineering and
Seismology (2ECEES), Istanbul, Turkey, paper 686.
Mwafy, A., Elnashai, A., & Yen, W.-H. (2007). Implications of
design assumptions on capacity estimates and demand
predictions of multispan curved bridges. Journal of Bridge
Engineering, 12, 710 726.
Naeim, F., & Kelly, J.M. (1999). Design of seismic isolated
structures, from theory to practice. New York, NY: Wiley.
National Institute for Standards and Technology [NIST] (1996).
The January 17, 1995 Hyogoken-Nanbu (Kobe) earthquake:
Performance of structures lifelines, and fire protection
systems, NIST Special Publication 901 Author.
Poulos, H.G., & Davis, E.H. (1980). Pile foundation analysis and
design. New York, NY: Wiley.
Priestley, M.J.N., Seible, F., & Calvi, G.M. (1996). Seismic
design and retrofit of bridges. New York, NY: Wiley.
Public Works Research Institute [PWRI] (2011). Technical note
of PWRI no. 4217. In Proceedings of the 43rd Joint Meeting
of U.S.-Japan panel on wind and seismic effects (UJNR),
Tsukuba, August 29 30 (pp. 154 163). Tsukuba: Author.
Retrieved from http://www.nehrp.gov/pdf/UJNR-4217.pdf
Sextos, A., Kappos, A., & Pitilakis, K. (2003). Inelastic dynamic
analysis of RC bridges accounting for spatial variability of
ground motion, site effects and soil-structure interaction
phenomena. Part 2: Parametric analysis. Earthquake
Engineering and Structural Dynamics, 32, 629 652.
Stanton, J.F., Roeder, C.W., Mackenzie-Helnwein, P., White, C.,
Kuester, C., & Craig, B. (2008). Rotation limits for
elastomeric bearing (NCHRP Report 596). Washington,
DC: National Cooperative Highway Research Program,
Transportation Research Board, National Research Council.
Yang, Q.R., Liu, W.G., He, W.F., & Feng, D.M. (2010). Tensile
stiffness and deformation model of rubber isolators in tension
and tension-shear states. Journal of Engineering Mechanics,
136, 429 437.

También podría gustarte