You are on page 1of 16

This article was downloaded by:[Sivakumar Babu, G.L.

]
On: 24 May 2008
Access Details: [subscription number 793268158]
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

Geomechanics and Geoengineering


An International Journal
Publication details, including instructions for authors and subscription information:
http://www.informaworld.com/smpp/title~content=t725304177
Buckling and bending response of slender piles in
liquefiable soils during earthquakes
Sumanta Haldar a; G.L. Sivakumar Babu a; Subhamoy Bhattacharya b
a
Department of Civil Engineering, Indian Institute of Science, Bangalore, India
b
University of Oxford, Oxford, UK

Online Publication Date: 01 June 2008


To cite this Article: Haldar, Sumanta, Sivakumar Babu, G.L. and Bhattacharya,
Subhamoy (2008) 'Buckling and bending response of slender piles in liquefiable
soils during earthquakes', Geomechanics and Geoengineering, 3:2, 129 — 143
To link to this article: DOI: 10.1080/17486020802087101
URL: http://dx.doi.org/10.1080/17486020802087101

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf


This article maybe used for research, teaching and private study purposes. Any substantial or systematic reproduction,
re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expressly
forbidden.
The publisher does not give any warranty express or implied or make any representation that the contents will be
complete or accurate or up to date. The accuracy of any instructions, formulae and drug doses should be
independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings,
demand or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or
arising out of the use of this material.
Geomechanics and Geoengineering: An International Journal
Vol. 3, No. 2, June 2008, 129--143
Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

Buckling and bending response of slender piles in liquefiable soils during earthquakes
Sumanta Haldara, G.L. Sivakumar Babua,* and Subhamoy Bhattacharyab
a
Department of Civil Engineering, Indian Institute of Science, Bangalore 560 012, India; bUniversity of Oxford, Oxford OX1 3PJ, UK

(Received 07 June 2007; in final form 14 March 2008)

Design of pile foundations in seismically liquefiable soils involves identifying the appropriate failure mechanisms. Piles in liquefiable soils are
conventionally designed against bending failure due to lateral loads arising from inertia and/or lateral spreading. This is strong evidence that there is
another mechanism, which the code does not consider, that may govern the failure of these foundations. In this paper, the response of a single end
bearing pile in liquefied soil with and without the effect of axial load has been presented. The effect of liquefaction is incorporated in the pile--soil
interaction through nonlinear analysis using the finite difference program Fast Lagrangian Analysis of Continua (FLAC). The method of analysis is
carried out using the well documented failure of Showa Bridge piles which failed during the 1964 Niigata earthquake. The response of the pile is also
evaluated using dynamic analysis. The need for proper identification of failure mechanisms as well as design guidelines is highlighted.

Keywords: design codes; buckling; liquefaction; moment; pile failure

1. Introduction The current understanding of pile failure is based on a [2]


bending mechanism, where the lateral loads due to inertia and
The analysis and design of piled foundations in seismically slope movements (lateral spreading) induce bending failure in
liquefiable soil has drawn considerable attention in recent the pile (Figure 1(a)). The evidence of large ground displace-
years due to reported failure in some of the recent earthquakes ments in the vicinity of the piled building and the tilting of the
(Bhattacharya et al. 2004, 2005a, 2005b, Tokimatsu and structure towards the slope seem to support the hypothesis of
Boulanger 2005, Bhattacharya 2006, Shanker et al. 2007, this mechanism. These ground movements are more pro-
Tazoh 2007). Design guidelines have been developed such as nounced near the quay wall or a waterfront or in sloping
the IS 1893 (2002), Japanese Code of Practice (JRA, 1996), grounds. As a result, permanent lateral deformation or lateral
Eurocode 8: part 5 (1998), USA code (NEHRP, 2000). Pile spreading of soil has been reported to be the main source of
failures are still reported; for example the damage survey of pile distress in piles (Abdoun and Dobry 2002, Finn and Fujita
supported structures following the earthquakes of Kobe 2002). It has also been reported that such large displacements
(1995), Bhuj (2001) and Sumatra (2005). The response of are the root cause of many bridge failures. For example
pile foundations during earthquakes is a very complex phe- Hamada (1992a, 1992b), Ishihara (1997), Tokimatsu et al.
nomenon which involves inertial and kinematic interaction (1998), Goh and O’Rourke (1999), Abdoun and Dorby
between pile and soil. This interaction has considerable sig- (2002), Finn and Fujita (2002), Berrill and Yasuda (2002),
nificance in developing seismic-resistant design methodolo- Finn (2005) and Miwa et al. (2006) carried out similar studies
gies. This is strong evidence that the mechanisms underlying and reported that the observed pile failures are due to large
the codes of practice may not be adequate or the calculation displacement followed by lateral spreading of the ground.
procedures in the codes under-predict the forces and moments Tokimatsu and Suzuki (2004) suggested that the earth pres-
in the pile. This paper suggests that the codes of practice sure variations due to large relative displacement can cause
overlook the effect of axial load on the lateral stiffness of the the failure of piles. These observations have led the design
pile when the soil surrounding the pile is in liquefied condi- codes to include checks on the bending moments in piles
tion. This is essentially a buckling mechanism which may caused by lateral spreading of the ground. For an example,
govern the failure of most slender piles in liquefied soil. In the Japanese Code of Practice (JRA 1996) advises engineers to
other words, the interaction between buckling instability and design piles with respect to bending failure assuming that a
bending needs to be considered. non-liquefied crust applies its full passive earth pressure and
the liquefied soil 30% of the total overburden pressure to the
pile (Figure 1(b)).
Bhattacharya (2003) and Bhattacharya et al. (2004, 2005)
*Corresponding author. Email: gls@civil.iisc.ernet.in demonstrated that the current understanding of pile failure

ISSN 1748-6025 print=ISSN 1748-6033 online


 2008 Taylor & Francis
DOI: 10.1080=17486020802087101
http:==www.informaworld.com
130 S. Haldar et al.

analysis as this will decide the magnitude of kinematic


Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

(a)
forces in the pile and also the unsupported length of the
pile. The depth of liquefaction is usually calculated based
on empirical (Seed and Idriss 1971) or semi-empirical
(Idriss and Boulanger 2004) procedures. The objective of
FLOWING SOIL
the paper is therefore to carry out a comprehensive study of
an end-bearing single pile in a liquefied soil using numerical
Water Table Non-liquefied
stabilised crust
dynamic analysis where the interactions can be accounted
for. The existing design approaches can then be compared.
This would also enable the prediction of failure based on a
Liquefied particular failure mechanism and also provide a better
Depth of layer
liquefaction
understanding of the merits and limitations of the different
design codes of practice. The soil-pile interaction response
is modelled using the finite difference program Fast
Lagrangian Analysis of Continua (FLAC) (Itasca 2006).
The well-known failure of piles of Showa Bridge is studied
as a case example throughout the paper. The design impli-
cations of the study are highlighted.
(b)

non -liquefiable
layer 2. Review of current design methods of pile design
HNL qNL qNL = Passive earth pressure
It is of interest to review the international codes of practice of
qL qL = 30% of overburden pressure seismic pile design. The Eurocode 8 advises designers to
liquefiable layer HL
design piles against bending due to inertia and kinematic
forces arising from the deformation of the surrounding soil.
non-liquefiable layer The code suggests that piles shall be designed to remain
elastic. Because of large pile movement such as in liquefiable
Figure 1. (a): Schematic diagram of the current understanding of pile failure. ground, the sections of the potential plastic hingeing shall
(b). JRA (1996) guideline for seismic design of pile in liquefied soil. be assumed for a region of 2d from the pile cap and a region
of  2d from any interface between two layers with markedly
different shear stiffness (ratio of shear moduli . 6), where ‘d’
denotes the pile diameter.
based on lateral spreading of the soil cannot always explain the
The Japanese design specification for highway bridges
cause of collapse and showed that while the design of the
(JRA 1996 or 2002) reported that liquefaction-induced lateral
Showa Bridge would be considered safe based on the current
spreading was the main cause of pile failure in a liquefied
provision of the JRA code, the bridge actually collapsed during
soil. As a result, guidelines were introduced to take into
the 1964 Niigata earthquake. They argued that the cause is the
account the forces due to liquefaction-induced ground move-
incorrect simplification of the effect of axial loads on the piles.
ment. The code advises practising engineers to design piles
Basically, a pile would become laterally unstable under the
against bending failure, assuming that the non-liquefied crust
axial load alone, if it is unsupported due to liquefaction.
exerts passive pressure on the pile and the liquefied soil 30%
Lateral loads or imperfections would reduce the failure load
of total overburden pressure. The concept is presented in
but are not a necessary condition for failure. Therefore the
Figure 1(b).
slenderness ratio of the pile in liquefiable soil is quite impor-
There is no such guideline available in IS1893-Part 1 (2002)
tant. Essentially, part of the piles in a liquefied zone should be
and the need for such guidelines is felt after pile failure in Bhuj
treated as long slender columns carrying lateral loads and
(2001) earthquake due to liquefaction.
should not be analysed as a simple cantilever beam carrying
lateral loads. In this paper, two failure mechanisms are considered
in detail: (a) buckling instability predominantly due to axial load
effects; (b) bending failure due to the lateral loads. Evaluation of
the performance of a piled foundation against these failure 3. Pile--soil interactions during earthquakes
mechanisms would require a reliable estimation of liquefiable
depth of soil and also the slenderness ratio of the pile in liquefi- Pile foundations in liquefiable soils under earthquake excitation
able zone. The available methods for analysis of piled foundations are subjected to various mechanisms and processes. Some
rely on the evaluation of liquefaction depth and do not capture the of the mechanisms and processes can be grouped under the
soil-pile interaction. The depth of liquefaction is crucial in the following headings:
Geomechanics and Geoengineering 131

lateral displacement of the piers might have caused five simply


Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

Y supported spans to fall. The bridge was founded on nine tubular


piles having 0.609 m diameter and 25 m length. The vertical
steel piles were driven through a 10 m layer of loose medium
sand into a 6 m underlying layer of dense fine sand. The 10 m
P layer is believed to have liquefied from the level of the riv-
erbed down to its full depth (Hamada 1992, Ishihara 1993).
Hinge formation
From the shape of the deformed pile (Figure 4) of the bridge,
Berrill and Yasuda (2002) inferred that the movement of the
liquefied sand layer towards the centre of the river may have
caused the deformation of the pile. They also hypothesised
Resistance offered that the forces to dislodge the decks from the pier may not
by soil have been transmitted through the inertia of the superstructure
or from the abutments, as they themselves were displaced by
L
movement of the riverbank. However, Bhattacharya (2003)
and Bhattacharya et al. (2005) showed that the hypothesis of
the lateral flow of liquefied soil could not explain the collapse
of the Showa Bridge. He argued that the piers close to the river
X banks did not fail, wherein the lateral flow was seen to be
severe.
Figure 2. Failure due to pile buckling (Bhattacharya, 2003). This paper aims to analyse one of the end-bearing piles of
Showa Bridge The justification for selecting this case is that
piles in this case history are very well documented. The soil
profile is divided into three layers and parameters are sum-
1. Soil and pile response due to cyclic loading of the earth- marized in Table 1. Bhattacharya et al. (2005) estimated that
quake: because of generation of excess pore water pressure the average static axial load of the bridge super structure is
in the soil, the effective stress of the soil comes to near zero about 6660 kN. There are nine piles in a row sharing the load
value. The piled structure vibrates. of the superstructure (Figure 5). Therefore the load acting on
2. Pile response to soil stiffness degradation: if the unsup- the pile each pile is about 6660 kN / 9 = 740 kN. Details of
ported length of the pile exceeds the length required for the pile section are summarized in the Table 2. The natural
the critical buckling length, it may become unstable and frequency of the bridge estimated by Fukuoka (1966) is about
buckle laterally in the direction of least elastic stiffness, 2 sec.
e.g. Figure 2 after Bhattacharya (2003). A reanalysis of the pile in Showa Bridge may give insight
3. Soil response to stiffness degradation and the resulting into the design issues of pile foundation in liquefiable soil. An
pile behaviour: in this phase, lateral flow of the liquefied overview of the methods stipulated in various design codes
soil may commence. Any non-liquefied crust at the top of used in practice gives possible advantages and limitations of
the liquefied soil may also flow with the liquefied soil. the different codal provisions. Hence an attempt is made to
This would load the pile severely. If, on the other hand, examine the design methodologies of pile, stipulated in differ-
the pile becomes unstable, it may shear the liquefied soil ent codes of practice using the present case example. The
in front of it. response is also analysed using dynamic numerical analysis,
and design implications for piles in liquefiable soil are
suggested.

4. A case study: Showa Bridge


4.1 Response of pile as per seismic design codes
Several case histories have been documented where the piled
foundations either survived or were severely damaged during Design provisions stipulated in design codes, namely
the earthquakes; see for example the quantitative analysis of Indian Standard, IS: 1893, Part 1 (2002), Eurocode 8,
fourteen case histories by Bhattacharya (2003), Bhattacharya Part 5 (1998), JRA (1996), are described in the following
et al. (2004), and three case histories by Lin et al. (2005). sections. The details of pile and soil properties of the case
In some cases, large amount of lateral ground displacement example are presented in Tables 1 and 2 respectively.
was observed, for example the Niigata earthquake (1964). IS: 1893 and Eurocode provide calculation procedures for
Among all the failures, the most striking failures are the col- inertial bending effects. On the other hand, JRA (1996)
lapse of Showa Bridge (Figure 3). Details of the collapse are provides calculation procedures for kinematic effects.
documented in the literature: Hamada (1992), Ishihara (1993), Essentially, all the codes check the safety against linear
Bhattacharya et al. (2005). Hamada (1992) reported that the static bending.
132 S. Haldar et al.
Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

Figure 3. Damage of pile in Niigata Earthquake, 1964 (Finn et al., 2002).

4.1.1 IS: 1893-Part 1 (2002) and IS 2911 (1997) structure. The approximate fundamental natural period of
In India, two codes are used for seismic design of piles. They are: vibration (Ta), in seconds is considered as 2 sec. The soil in
the site can be classified as soft soil, based on I.S:1893 (2002)
1. IS: 1893 (2002), which is the code of practice for estimat- and spectral accelerations for different vibration periods for soft
ing the forces on the superstructure as well as the base soil can be calculated as below.
shear.
2. IS: 2911 (1997), the code of practice for pile design.   1 þ 15Ta ; 0:0 # Ta # 0:10
Sa
¼ 2:5; 0:10 # Ta # 0:67 ð3Þ
g
1:67=Ta ; 0:67 # Ta # 4:0

4.1.1.1 IS: 1893-Part 1 based design where Ta = vibration period of the structure.
The seismic base shear acting on the pile is usually estimated, The site location is assumed to be situated in severe earth-
for example IS: 1893 (or Eurocode: 8), by the equivalent lateral quake zone and hence a value of Z = 0.36 is selected. The value
force procedure given by, of the importance factor (I) is assigned as 1.5, following IS Code
as the structure is a bridge (lifeline structure). The value of R is
VB ¼  W ð1Þ selected as 3 considering the ordinary moment resistance of the
bridge structure. The value of Sga ¼ 0:835 for Ta = 2 s;
where  = horizontal seismic co-efficient, W = total dead The calculated seismic co-efficient  is 0.075.
load from superstructure. Total dead load on each pile is computed as 740 kN (Table 2)
and hence the base shear force on each pile can be estimated by,
 
Z:Ifactor Sa
¼ ð2Þ
2R g VB ¼ 0:075 · 740 kN ¼ 55:6 kN ð4Þ

where Z = zone factor for the maximum considered earth-


quake (MCE), Ifactor = importance factor for the structure,
R = response reduction factor, depending on the perceived 4.1.1.2 Estimating maximum moment in the pile based on pile
seismic damage performance characterized by ductile or brittle design code IS: 2911
deformations, Sga = average response acceleration coefficient, The underlying mechanism behind the IS: 2911 is the approx-
which is a function of the site and vibration period of the imate method of Davisson and Robinson (1965) to analyse
Geomechanics and Geoengineering 133
Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

Total dead load = 6660 kN

Bridge deck

9 piles in a row

Figure 5. Schematic diagram of the Showa Bridge deck and piles.

Table 2. Pile properties

Pile parameters Values

Length, (m) 25
Outer diameter, Do(m) 0.609
Inner diameter, Di (m) [above ground level, 0 to + 9 m] 0.577
Inner diameter, Di (m) [0 to -3 m] 0.577
Inner diameter, Di (m) [-3 to -16 m] 0.591
Moment of inertia, I (m4) [above ground level, 0 to + 9 m] 1.31 · 10-3
Moment of inertia, I (m4) [0 to -3 m] 1.31 · 10-3
Moment of inertia, I (m4) [-3 to -16 m] 7.64 · 10-4
Elastic modulus, E (N/m2) 2.10 · 1011
Vertical load, P (N) (Bhattacharya 2003) 0.74 · 106
Material Steel

sffiffiffiffiffiffi
5 EI
T¼ ð5Þ
h

where EI = stiffness of the pile, hh = modulus of subgrade


reaction having units of force/ length3 and typical values are
given in Table 3. The point of fixity in this case is calculated as
follows:
Figure 4. Failed pile and the soil profile, Fukuoka (1966).
The average SPT-N value at the top layer is 4 (Table 1), and
therefore the relative density can be estimated by Equation (6),
given by Meyerhof (1957). Meyerhof (1957) proposed an
Table 1. Soil properties
empirical correlation between SPT-N, effective vertical stress
Parameters Soil 1 Soil 2 Soil 3 and relative density (R.D in %), as follows:
Depth (m) 0--3 m 3--10 m 10--30 m sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
(N1)60 4 7 33 N
Shear modulus, G0 (Pa) 3.99·107 5.84·107 1.68·108 R:Dð%Þ ¼ 21 ð6Þ
Bulk modulus, K (Pa) 1.86·108 2.73·108 7.83·108 0v ðkgf=cm2 Þ þ 0:7
Density (kg/m3) 2000 2000 2000
Shear wave velocity, 141 170 289
Vs (m/sec) where N = SPT-N value, v¢ = vertical effective stress in
Dr (%) 30 40 86 kgf/cm2. For a SPT-N value of 4, the relative density is esti-
Angle of internal friction,  28 29 37
Poisson’s ratio, n 0.4 0.4 0.4 mated to be about 42% and therefore the modulus of subgrade
Constitutive model Byrne (1991) Byrne (1991) Mohr-Coulomb
Table 3. Values of the modulus of subgrade reaction

Value of hh (MN/m3)
laterally loaded pile. In this procedure, a laterally loaded pile is Relative density Sand below the water table Sand above the water table
assumed to be fixed at some point in the ground, the depth
of which depends on the relative stiffness between the soil and 40% 8 13
60% 24 42
the pile. This method, widely used in practice, involves the 80% 40 75
computation of stiffness factor T defined by,
134 S. Haldar et al.

reaction is 10 MN/m3 and value of EI=160.65 MNm2. The where Se(Ta) is the ordinate of elastic response spectrum, Ta is
Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

stiffness factor T, defined by Equation (7), can be calculated by, the vibration period of a linear single degree of freedom system,
rffiffiffiffiffiffiffiffiffiffiffiffiffiffi ag is the design ground acceleration for the reference return
5 160:65 period, 0 is the spectral acceleration amplification factor for
T¼ ¼ 1:74 m ð7Þ 5% viscous damping, TB, TC are the limits of the constant
10
spectral acceleration branch, TD is the value defining the begin-
Therefore for L1/T = 9/1.74 = 5.74 (L1 is the projected ning of the constant displacement range of spectrum, k1, k2 are the
length of pile above ground level, here L1=9 m) the point of exponents which influence the shape of the spectrum for a vibra-
fixity is at 1.8 · 1.74 = 3.1 m (Lf/T = 1.8, Lf is the point of tion period greater than TC, TD respectively, S is the soil para-
fixity). The moment in the foundation can be estimated by, meter, g. is the gravitational constant, q is the behaviour factor
given for various structures and ductility levels. The value of the
MR ¼ 55:6 kN · ð3:1 þ 9Þ m ¼ 672:76 kNm ð8Þ above parameters for the soil class C is given in Eurocode 8:
S = 0.9;  0 = 2.5; q = 4; k1 = 1.0; k2 = 2.0; TB = 0.2;
The
 plastic moment
 capacity of the pile section (9 mm thick) TC = 0.8; TD = 3.0; and ag = 2.18 m/s2 (1964 Niigata
¼ 0:6093 0:5913
m · 500 · 103 kPa ¼ 1620 kNm.
3 earthquake)
6  6
At Ta = 2 s, design spectral acceleration, Se/g = 2.18/9.81 ·
In summary, the Indian code would predict a moment 672.76 0.9 · 2.5/4 · (0.8/2)1.0=0.05
kNm. The plastic moment carrying capacity of Showa Bridge
The base shear VB= 0.05 · 740 = 37 kN
pile is about 1620 kNm and therefore the bridge should be
considered safe according to I.S. codes.
MR ¼ 37 kN · ð3:1 þ 9Þ m ¼ 447:7 kNm ð10Þ
4.1.2 Eurocode 8: Part 5 (1998)
The soil type in the present case follows the Class C soil in
Eurocode 8. The design spectrum acceleration values can be
calculated as: 4.1.3 JRA (1996)
Figure 6 shows the loading diagram based on the JRA code
h  i
ag for the lateral spreading mechanism. JRA is the only code
g S 1 þ TTBa q0  1 ; 0 # Ta # TB
of practice that advises a calculation procedure for evaluat-
ag S q0 ; TB # Ta # TC ing the response against lateral spreading. The calculation
Se ðTa Þ ¼ h ik1 ð9Þ below estimates the maximum moment. Inertia of the super-
ag S q0 TC
Ta ; TC # Ta # TD
h ik1 h ik2 structure can also induce bending moments in the pile. The
ag S q0 TC
TD
TD
Ta ; TD # Ta JRA code advises engineers not to superimpose the effects

Water level 9m
Mud line
3m (assumed)

9kPa (based A
on JRA 1996)

69kPa (based 10m


on JRA 1996)

Dense non- 6m
liquefiable
soil

Figure 6. Schematic diagram showing the predicted loading based on JRA code.
Geomechanics and Geoengineering 135

of lateral spreading and inertia. The reason behind such prag- Experiments show that the actual failure load of slender
Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

matic advice is explained in Ishihara (1997). He notes that the columns is much lower than the critical buckling load predicted
soil liquefaction starts during an earthquake at approximately by Equation (11). Rankine (1866) recognised that the actual
the instant of peak acceleration. He argues that since the failure involved an interaction between elastic and plastic
seismic motion has already passed its peak, such shaking as modes of failure. Lateral loads or inevitable geometrical imper-
may persist will be less intense, so that the inertia force fections lead to creation of bending moments in addition to
transmitted to the superstructure will not be significant. axial loads. Bending moments have to be accompanied by
stress resultants that diminish the cross-sectional area available
for carrying the axial load, so the failure loads PFailure , Pcr.
4.1.3.1 Calculations based on JRA (1996) Equally, the growth of zones of plastic bending reduces the
Assuming the bulk unit weight of soil is 20 kN/m3, maximum effective elastic modulus of the section, thereby reducing the
lateral spreading pressure at mudline at point A in Figure critical load for buckling. As the elastic critical load is
6 = 30% of total overburden pressure due to water = 0.3 · approached, all bending effects are magnified. Timoshenko
10 kN/m3 · 3 m = 9 kPa. Maximum lateral spreading pressure and Gere (1961) showed that in the case of stability analysis
at 10 m depth acting at point B in Figure 6 = 30% of total of elastic columns, the lateral displacement caused by lateral
overburden pressure = 0.3 · (20 kN/m3 · 10 m + 10 kN/m3 · loads is greatly amplified in the presence of axial loads. If 0 is
3 m) 69 kPa. Maximum moment, at point B in Figure 6, due to the displacement of a cantilever column due to lateral loads
spreading force (trapezoidal loading) = (0.5 · 60 kPa · 10 m · alone, the final displacement () gets amplified in the presence
0.609 m · 3.33 m) + (9 kPa ·10 m · 0.609 m · 5 m) = 608 of axial load (P) by the following expression:
kNm + 274 kNm = 882 kNm. The predicted moment is much
less than the plastic moment capacity of the pile and therefore the 1  1
JRA (1996) would consider the bridge to be safe.  ¼ 0   or; ¼  ð12Þ
1  PPcr 0 1  PPcr

4.1.4 Summary of the analysis based on the various codes of  


The term 0 can be termed as the ‘‘Buckling Amplification
practice
Factor’’ and is amplification of lateral displacements due to the
Calculations based on IS-1893 (2002) (Indian Standard),
presence of axial load. Figure 7 presents a graph of buckling
Eurocode 8: Part 5 (1998) and JRA indicate that the predicted
amplification factor plotted against the normalized axial load
bending moment is less than the plastic moment carrying capa-
ðPPcr Þ where P denotes the applied axial load. It can be observed
city of pile section (1620 kNm) and therefore the pile could be
from the graph and Equation (12) that if the applied load is 50%
considered safe in bending. However, the bridge failed during
of Pcr, the amplification of lateral deflection due to lateral loads
the 1964 Niigata earthquake and therefore it may be concluded
is about 2 times. At these large deflections, secondary moments
that some other governing mechanism has not been considered.
generate, which leads to more deflections and more bending
In other words, the methodologies in the different design codes
moment. It would therefore be important to be in the linear
are simple, but do not necessarily capture the actual failure
regime and not in any way near the asymptotic region. It would
mechanism.

4.2 Failure mechanism based on buckling instability 10

The possible failure mechanism of the Showa bridge pile as 9


described by Bhattacharya (2003) is due to the liquefaction of 8
Amplification of displacement

the top soil layer and the stiffness of the soil being reduced to
7
near zero value. Therefore, the pile behaves like an unsupported
column and it can become unstable and it moves to a position of 6
no return -- instability type failure. Euler’s critical buckling load
Present case (P/Pc r=0.82)

5
of the pile is defined by:
No buckling Intiation of buckling
4
2

Pcr ¼ EI ð11Þ 3
L2eff
2

Leff is the effective length of the column, which depends on the 1


boundary conditions of the column ends. The critical load (Pcr)
0
of an axially loaded structure is essentially the minimum axial 0 0.2 0.4 0.6 0.8 1
load at which an initially straight stress-free column becomes P/Pcr
unstable and the transverse deflection becomes indefinitely
large. Figure 7. Amplification of lateral displacement versus normalised axial load.
136 S. Haldar et al.

also be unwise to use a factor of safety less than 3 against the analysis to a limited extent. This type of analysis may conclude
Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

Euler load of a pile, i.e. ðPPcr ¼ 0:33Þ. In such cases, there are no that a pile is unsafe against bending mode or buckling instability
chances of amplification of lateral deflections due to axial loads but will not provide any information about the dominant mode at
and the point lies in the linear range of Figure 7. Structural the intermediate stages of liquefaction, i.e. at partial liquefaction
engineers generally prefer to provide a factor of safety of at stages, which can be obtained from dynamic numerical analysis.
least 3 against linear elastic buckling to take into account the Numerical analysis also enables proper consideration of an appro-
eccentricity of load, deterioration of elastic stiffness due to priate constitutive model for soil and pile--soil interactions under
plastic yielding and unavoidable imperfections. The actual fail- dynamic conditions. Hence, in the following section the numer-
ure load (Pfailure) is therefore a factor, ( , 1) times the ical analysis is conducted for one of the Showa Bridge piles.
theoretical Euler’s buckling load given by Equation (13).

Pfailure ¼ : Pcr ð13Þ 5. Method of analysis

Physical modelling, numerical modelling or study of field case


Hence, it may be inferred that instability may be expected at records is generally adopted to determine the behaviour of piles
around 0.35, i.e. is taken as 0.35. However, this factor will under earthquake excitation. Physical models are sometimes
depend on the axial load, imperfections or the residual stresses very expensive, and numerical models predict better results if
in the pile due to driving. proper soil--pile models are considered. Coupled analysis, i.e.
In the present case the length of the pile that is most likely to incorporating the fluid--soil--structure interaction, demands not
be unsupported during liquefaction is the sum of zone of lique- only high computational effort but also proper understanding of
fied soil (10 m) and the depth of fixity below the liquefiable the interactions. Therefore, various simplified approaches are
soils (typically 2 m). On adding the length of pile in free air/ used in practice, namely the Winkler spring and dashpot system,
water of 9 m, the unsupported length is 21 m. From the buckled p-y approach, pseudo-static approach, etc. Finn and Fujita (2002)
shape (shown as original position in Figure 4), it is clear that the report that dynamic non-linear finite element continuum analysis
pile was in a fixed--free boundary condition and hence the gives very accurate results. Several researchers performed
effective length is twice the length in unsupported zone. numerical studies on piles under earthquake loading using lin-
ear/non-linear dynamic finite element analysis. Kaynia and
Kausel (1991) provided elastic solutions for pile groups based
2 · 160:62 · 103 kNm2 on boundary integral techniques. Wu and Finn (1997a, 1997b)
Pcr ¼ ¼ 898 kN ð14Þ
ð2ð9 þ 12ÞÞ2 m2 studied the pile response by quasi-3D dynamic analysis.
Liyanapathirana and Poulos (2005) observed the lateral response
of the pile in liquefying soil using a non-linear Winkler model
Therefore the ratio of PPcr ¼ 740
898 ¼ 0:82. The lateral displace- and results were validated with centrifuge test results. Arduino
ment amplification factor described by equation (12) is plotted et al. (2005) performed the numerical analysis of dynamic stiff-
and presented in Figure 7. The calculated PPcr value for the pre- ness of the pile in liquefiable soil. From the above discussion it is
sent analysis is also shown in the same figure. From Figure 7, it clear that numerical analysis is a useful tool for the analysis of
is evident that if the applied axial load (P) is 82% of the critical pile response in liquefied soil. The present study focuses on the
buckling load (Pcr), the lateral displacement is amplified by pile response based on numerical analysis by FLAC. The results
about 6 times due to the application of lateral load. In other based on design codes indicated earlier are examined in relation
words, at the instant of liquefaction, due to the presence of axial to the results obtained from dynamic numerical analysis and the
load, the pile deflects laterally and cannot move back to equili- design implications are highlighted.
brium position. From Figure 7, it can be noted that in the PPcr
regime there is little chance of amplification of the lateral
deflections due to axial loads. Hence the allowable load to 5.1 Dynamic liquefaction analysis by explicit finite
avoid buckling instability-type failure is about 898/3  300 difference method
kN. It must be remembered that the applied dead load on each
pile was 740 kN, which is much higher than the allowable load FLAC analysis permits two-dimensional, plain strain analysis.
against buckling considerations. The calculation is based on an explicit finite difference
The above simple calculation shows that the buckling instabil- scheme to solve the equation of motion using lumped grid
ity might be the cause of bridge failure. One of the main uncer- point masses derived from the real density of surrounding
tainties in the buckling calculations is the unsupported length of zones. The FLAC formulation is coupled to the structural
the pile during liquefaction. The current methods of evaluation of element model, thus permitting analysis of soil--structure
liquefaction at the given site can provide us with the most likely interaction brought about by ground shaking. Dynamic soil--
depth of liquefaction depending on parameters such as the PGA structure interaction is captured using the hysteresis curves
(peak ground acceleration) and moment magnitude of earthquake and energy-absorbing characteristics of real soil. The dynamic
(Idriss and Boulanger 2004). These depths are helpful for failure loading is applied in terms of input boundary condition at the
Geomechanics and Geoengineering 137

model boundary. In loose sands pore pressure may build up the superstructure will impose dynamic axial loads on the piles
Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

considerably due to cyclic shear loading. Eventually by this which increase the axial load on some piles. This increase may
process effective stress approaches zero. FLAC has built in range from 10% to as high as 50% depending on various factors
constitutive models, namely the Byrne model (Byrne 1991) for such as the type of superstructure, height of the centre of mass of
pore pressure generation. The static constitutive models, e.g. the the superstructure, frequency and magnitude of the earthquake.
Mohr-Coulomb model, can also be used by adding on to it the This factor is termed the ‘dynamic axial load factor’ denoted by
required Rayleigh damping, geometric damping and modulus  (Bhattacharya 2006). The following equation gives an estimate
reduction curves. The Byrne models considered soil behaviour of the maximum axial load acting on a pile.
due to cyclic loading such as energy dissipation, volume changes
and modulus degradation. The present study utilizes the Byrne
Pdynamic ¼ ð1 þ ÞPstatic ð17Þ
model for modelling the soil. The constants in the constitutive
model are related to macro-scale soil property and the SPT-N
value of the soil. The constitutive model as proposed by Byrne In the case of the Showa Bridge failure, the bridge failed
(1991) is a simplification of the Finn model and states that at about 60 to 70 seconds based on the available report (e.g.
Hamada 1992), hence the present analysis considers  as zero.
  
"vd "vd
¼ C1 exp C2 ð15Þ
5.2.1 Boundary condition of the piles at the pile head
The superstructure of the Showa Bridge, i.e. the bridge decks,
where "vd is the incremental volume strain and "vd is were composed of panels alternately resting on roller and fixed
accumulated volume strain,g is the cyclic shear strain ampli- supports (Figure 8). The construction of the bridge was such
tude and C1, C2 are constants. Byrne (1991) describes the that one end of the girders was locked and the other end was free
relation of constants as C2 = 0.4/C1 and also with the other to slide longitudinally off the piers. For the purpose of analysis,
parameters as follows: it is considered that the lateral loads were large enough at peak
inertia to overcome the roller friction and therefore there was no
restraints on the movement of the piles. To take into account the
C1 ¼ 7600ðDr Þ2:5 ; Dr ¼ ðN1 Þ0:5
60 or; inertial effect, a pseudo-static lateral inertial load of 10% of the
ð16Þ
C1 ¼ 8:7ðN1 Þ1:25
60
axial load which is typical of such bridge is adopted. In order
words, 74 kN of lateral load is applied at the pile head. A
rotational restraint is applied at the pile head so that it cannot
(N1)60 is the normalized SPT-N value with respect to over-
rotate but it can deflect laterally.
burden pressure of 100 kPa and corrected to a ratio of 60%.
Once the liquefaction starts, i.e. after the peak inertia is
Dr is the relative density. The modelling assumptions and
reached, the pile head undergoes a severe displacement and
details for FLAC analysis in the present study are summarised
resistance offered from the bridge deck is minimized and the
in the following section.
pile head acts similarly to a free head. A boundary condition at
the pile head is imposed similar to a free-head pile after liquefac-
5.2 Study of the failure of showa bridge tion starts. The program uses the 4-noded quadrilateral grids.
Total soil medium is discretised into 760 numbers of grids in 19
Soil is modelled using Mohr-Coulomb failure criteria and the rows and 40 columns. In the lateral direction each grid side
pilesis modelled as a -wo noded pile element. Donovan et al. dimension is 1.5 m. For 0--3 m depth (Soil 1) the grid dimension
(1984) recommend linear scaling of material properties to in the vertical direction is 0.5 m, 3--10 m (Soil 2) is 1.4 m and 10--
distribute the discrete effect of elements over the distance 30 m (Soil 3) is 2.2 m respectively. The pile--soil shear stiffness
between elements in a regularly spaced pattern. Hence, piles are is assigned in the FLAC model as 10-times stiffer than the stiffest
modelled as pile elements and appropriate soil--pile interface
properties, namely axial stiffness, normal stiffness and shear
stiffness are scaled to represent plane strain conditions using the
above procedure. Size of the field is considered as 60 m · 30 m.
A hollow steel pile length of 25 m and outer diameter of 0.609 m
is considered for the analysis and divided into 40 equal segments.
Bhattacharya et al. (2005) estimated that the average static axial
load acting on the pile is about 740 kN. Therefore, a constant
vertical load of 740 kN is applied to the pile.
The dynamic component of the axial load has not been
considered. In general, the static axial load (Pstatic) acts on
each pile beneath the structure, assuming that each pile is
equally loaded during static conditions, ignoring any eccentri- Figure 8. Schematic diagram of the Showa Bridge showing the restraint of the
city of loading. However, during earthquakes, inertial action of pile head.
138 S. Haldar et al.

neighbouring zone, as suggested by Comodromos et al. (2003). mesh size for the FLAC model is selected to ensure accurate wave
Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

Figure 9 represents the schematic diagram of soil-pile system transmission. Based upon the elastic properties listed in Table 1,
considered in the analysis. the least shear wave velocity is 141 m/s and largest zone size (l)
Dynamic characteristics of the soils in this model are assumed is 2.2 m. In order to provide reasonable runtime, the maximum
to be governed by the modulus reduction factor (G/G0). Wave frequency that can be modelled accurately is (Itasca 2006):
reflections at model boundaries are minimised by specifying free-
field boundary conditions. The dynamic loading boundary con- Vs 141
f¼ ¼  6:4 Hz ð18Þ
dition is applied as an acceleration history. A coupled dynamic 10 l 10 · 2:2
analysis including the effect of ground water is incorporated as
the soil is susceptible to liquefaction. The accelerogram data for
the Niigata site is obtained from Kudo et al. (2000). However, the Therefore the present mesh size will not affect the result if
data reported at the ground level have attenuated acceleration the input acceleration is filtered by removing the frequencies
values. Therefore some modification is necessary to determine above 6 Hz. Hence to save the dynamic analysis time and
the base input motion. Appropriately reduced base input motion is memory, the input acceleration is filtered to remove the
applied at base level and ground level acceleration is measured to frequencies above 6 Hz and the filtered acceleration is used
match with the data available as follows. Several trials are made for the present analysis.
to match the ground level acceleration. A scaling factor of 0.5 is
obtained and applied to available accelerogram data and used as
base input motion. The acceleration time histories at the ground 6. Results and discussion
surface as well as at the soil base are shown in Figure 10. The
6.1 Liquefaction analysis and ground response
P Figure 11 represents excess pore pressure ratio (EPWPR) with
Hi time. The excess pore water pressure build-up to liquefaction is
9m at 10 sec. The zone of liquefaction (hL) with depth is calculated
0m
Soil 1
–3m
Soil 2 1
EPWPR

–10m 0.75 Depth –1.0 m


0.5
6m
Soil 3 0.25
0
0 5 10 15 20 25
1
0.75 Depth –3.0 m
EPWPR

–30m 0.5
0.25
60m 0
0 5 10 15 20 25
Figure 9. Schematic diagram of pile and soil profile considered in the analysis.
1
0.75 Depth –4.0 m
EPWPR

0.5
3 0.25
0
Acceleration (m/sec2)

2 0 5 10 15 20 25
Ground level
1 1
0.75 Depth –6.0 m
EPWPR

0
0.5
–1 0.25
–2 0
0 5 10 15 20 25
–3
1
3 0.75 Depth –8.0 m
EPWPR
Acceleration (m/sec2)

2 0.5
Base level 0.25
1 0
0 0 5 10 15 20 25

–1 1
0.75
EPWPR

–2 0.5 Depth –9.0 m


–3 0.25
0 10 20 30 40 50 0
Time (sec) 0 5 10 15 20 25
Time (sec)
Figure 10. Acceleration time history of Niigata (1964) earthquake at the
surface and base level. Figure 11. Excess pore water pressure ratio at different depths.
Geomechanics and Geoengineering 139
Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

EPWPR EPWPR
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
0

Zone of liquefaction
4

Depth (m) 6

10
5 sec 10 sec
12

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1


0

4 Zone of liquefaction Zone of liquefaction


Depth (m)

10
15 sec 20 sec
12

Figure 12. Zone of liquefaction at different time instant.

Shear strain
as the Ru value greater or equal to 1.0. Hence, the zone of 0 0.025 0.05
liquefaction can be identified where Ru 1.0. Figure 12 shows 0
the zone of liquefaction at different times. As the top 10 m
layer is prone to liquefy, the calculation of Ru values is 2
restricted to a 10 m soil layer. The EPWPR is greater than or
equal to 1.0 for a depth of 7 m at 10 sec excitation and 8 m at 15
sec of excitation and hence it is classified as the zone of 4
liquefaction. The ground displacement is captured at different
Depth (m)

depths of soil by utilising the in-built history command in 6


FLAC. At each FLAC grid point the depth of soil is monitored
with respect to time of excitation. Large ground displacement
(up to 2 m) is observed at the site. 8
Figure 13 represents the development of shear strain within
the liquefied zone. The shear strain is increased up to 2% at 10
GL--4 m and 4.8% at GL--8 m. Below GL--10 m the SPT-N Time = 8 sec
values are relatively high. Therefore a small amount of shear Time = 15 sec
strain (0.5%) is observed. Hence it is considered that up to the 12
GL--10 m layer is liquefied due to earthquake. Figure 13. Shear strain in the liquefied zone.

6.2 Response of the pile considering axial load effects monitored with respect to its base displacement. Figure 14
represents relative lateral displacement of the pile due to the
The relative lateral displacement of the pile at different levels presence of vertical and lateral inertial load. The distribution of
along its length is observed. The lateral displacements are bending moment along the length of the pile is presented in
140 S. Haldar et al.

carrying capacity. Clearly, this is instability and at this large


Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

2 D : Pile head (+9 m from ground level)


deflection the pile will form a plastic hinge.
Relative lateral displacement of pile

D
C : At –9 m (At ground level)
1.5
B : At –12 m (–3m from ground level)

A : At –16 m (–7m from ground level) C


6.3 Interpretation of failure modes
1
Just after
(m)

Pile base Before liquefaction In this paper, two types of mechanism are considered: buckling
onset of B
0.5 liquefaction
(long column buckling) and excessive bending (beam bending).
A There is a subtle difference between bending and buckling
though in both the cases the pile will eventually fail by forma-
0
0 5 10 15 20 25
tion of a plastic hinge, which means the bending moment is
being exceeded. The essential difference between the two
–0.5 approaches can be very well understood from the structural
Time(sec)
design of columns and beams. For column design, it is checked
Figure 14. Relative lateral displacement of pile with respect to pile base. whether the column is ‘short’ or ‘long/ slender’. This is essentially
due to the fact that a long column will fail by buckling instability
and therefore we are interested in the stiffness (EI). In steel
Bending moment (kN-m) design, the above is carried out by estimating the slenderness
–500 0 500 1000 1500 2000 ratio of the column and then reducing the allowable stress based
0 on the Rankine and Merchant formula or the Perry-Robertson
5 sec formula (Timoshenko and Gere 1964). Buckling is very sensitive
10 sec to imperfections such as out-of-line straightness or initial curva-
ture, eccentricity of loading, or lateral loads due to the amplifi-
Plastic moment carrying capacity

5 15 sec
20 sec cation of lateral deflections. Practically, the failure due to
buckling will occur at axial loads which are much less than
Pile length (m)

10 the Euler elastic critical load, Pcr. As the elastic buckling


load of a pile is a function of the unsupported length, it
immediately becomes obvious that the depth to which the
15 soil would liquefy is an important factor.
On the other hand, in the hypothesis of pile failure by lateral
spreading, the bending moment induced in a pile is a function of
20 the kinematic forces acting on the pile due to the flowing soil. It
is therefore important to note that the bending moment induced
is also a function of the depth of the liquefiable layer. The
25
failure due to bending occurs when the maximum bending
moment of the pile is greater than or equal to the plastic moment
Figure 15. Bending moment profile with depth. capacity of the pile section. The critical buckling load is
directly related to the depth of the liquefiable layer (hL) and
the depth of fixity below the liquefiable soils (= 2 m) which
Figure 15. The bending moment in the pile head during the
will be the unsupported length of pile. The present analysis
initial part of the earthquake, i.e. at about 5 s, is perhaps due to
shows the effective length of pile is 2 · (9+10) = 38 m (8 m is
the restraint offered by the bridge deck, i.e. the fixed end
the depth of the liquefiable layer, 2 m is the depth of pile fixity
moment. The magnitude of the moment will depend on the
below the liquefied layer and 9 m is the unsupported length of
accurate estimation of the stiffness offered by the bridge and
pile above ground level) and the critical buckling load is
is not the main focus of the paper. This analysis simply shows
that the piles are safe against inertia loading (moment is much
less than 1620 kNm) and that the pile could not have failed 2 · 160:62 · 103
Pcr ¼ ¼ 1098 kN ð19Þ
before the onset of liquefaction due to the combined effect of ð2ð9 þ 10ÞÞ2
inertia and axial load.
The analysis also predicts a lateral deflection of the pile head Therefore the value of PPcr ¼ 0:7. From Figure 7 it is observed
of about 0.4 m before the onset of liquefaction, i.e. at about that if the applied axial load (P) is 70% of the critical buckling
10 sec in the time history. At this displacement, the out-of-line load (Pcr), the lateral displacement is amplified at least 3.5-
straightness (P-) moment will act on the pile and it is possible times due to the application of lateral load. To identify the axial
that the pile will be out of the roller support. As the liquefaction load that initiates buckling of the pile, a range of vertical load
progresses, there is a sudden increase in lateral displacement (74--740 kN) and lateral load of magnitude 10% of the axial
and the analysis predicts 1.64 m of pile head displacement. Till load is applied at the pile head. The relative lateral displace-
about 20 s, the bending moment is within the plastic moment ment of the pile head after complete liquefaction is obtained
Geomechanics and Geoengineering 141

whereas empirical/semi-empirical approaches suggest a


Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

5
Inertial load = 10 % of vertical load depth of 10 m, which means that the pile unsupported
Relative lateral displacement (m)
length is less based on the numerical study. While the

From numerical analysis P/Pcr = 0.7


4
numerical model incorporates some aspects of pile--soil
interaction, it seems to under-predict the depth of
No buckling liquefaction.
3
2. In the liquefied layer, the shaft resistance of the pile is
Initiation of reduced to zero and the axial load-carrying capacity of
2 buckling
the pile decreases. An appropriate design length of pile is
to be chosen so that pile capacity in liquefied soil is more
1 than the applied load. A guideline on reduction of pile
capacity and selection of design length of pile should be
implemented so that it can resist vertical load during
0 liquefaction.
0 0.2 0.4 0.6 0.8 1
P/Pcr 3. It is necessary to design against both bending and buckling
mechanisms. In particular, slender piles should be avoided
Figure 16. Amplified relative lateral displacement of pile with respect to pile in liquefiable regions.
base versus normalised axial load.

and is presented in Figure 16. The vertical load is represented as


P
Pcr ratios. The buckling instability of the pile is clearly observed 8. Concluding remarks
in Figure 16, and it can be noted that the lateral displacement of
the pile increases monotonically just after liquefaction and
This paper presents the response of pile--soil interaction analy-
much before the axial load of 740 kN. This can be attributed
sis in liquefied soil with reference to failed steel piles of Showa
to the presence of lateral loads, which indicates that the pile
Bridge during the Niigata (1964) earthquake. The limitations of
becomes unstable as liquefaction starts. It can be noted from
the existing design codes are highlighted. It has been shown that
Figure 16 in the PPcr regime that there is little chance of ampli-
the current codes of practice incorrectly simplify the effect of
fication of the lateral deflections due to axial loads. Hence, it is
axial load that the pile continues to carry at all times. The failure
clear that numerical analysis confirms that buckling instability
mechanisms of piles are described and an interpretation is
controls the response of the pile and the value of the critical
presented for two failure modes, namely buckling instability
buckling load is higher as the unsupported length is less.
and bending. From the design point of view, a pile in liquefiable
The observation that the maximum bending moment decreases
soil should be designed such that it can withstand failure against
after liquefaction indicates that the dominant failure mode is
buckling and bending and their combination. The need for
buckling instability.
proper identification of failure mechanisms is highlighted and
suggestions for use in current design codes are also indicated.

7. Design implications Acknowledgement

Current methods of pile design under earthquake loading are The authors are grateful to Prof. Nozomu Yoshida for his valuable comments
and technical support in doing the analysis. The authors are also thankful to the
based on a bending mechanism considering the inertia or lateral anonymous reviewers for useful comments and constructive criticism which
spreading of soil inducing bending moments in the pile, and the have been very useful in revising the paper.
effects of axial load are ignored. Based on the numerical ana-
lysis, the failure process of the pile foundation in liquefied soil
is described and it is evident that different design codes are References
inadequate and therefore appropriate design guidelines need to
be proposed. Abdou, T., and Dobry, R. 2002 Evaluation of pile foundation response
Based on the results presented, for the design of pile founda- to lateral spreading. Soil Dynamics and Earthquake Engineering,
tions in liquefied soil, some suggestions are listed for imple- 22, 1051--1058.
Arduino, P., Kramer. S.L., Li P., and Horne, J.C. 2005. Stiffness of
mentation in the current codes of practice as follows:
piles in liquefiable soils. In proceedings of a Workshop for Seismic
Performance and Simulation of Pile Foundations in Liquefied and
1. The failure of piles either by buckling or bending signifi- Laterally Spreading Ground, GSP 145. Davis, CA: ASCE,
cantly depends upon the depth of the liquefiable layer. The 134--148.
present numerical analysis of the failure of the Showa Berrill, J., and Yasuda, S. 2002. Liquefaction and piled foundations:
Bridge based on the Byrne constitutive model suggest Some issues. Journal of Earthquake Engineering, 6, Special Issue
that the top 8 m layer is liquefiable for the case analysed, 1.1--41.
142 S. Haldar et al.

Bhattacharya, S. 2003. Pile instability during earthquake liquefaction. Idriss, I.M., and Boulanger, R.W. 2004. Semi-empirical procedures for
Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

PhD thesis, University of Cambridge. evaluating liquefaction potential during earthquakes. Proc. 11th
Bhattacharya, S. 2006. Safety assessment of existing piled foundations Int. Conf. on Soil dyn. and Earth. Engg. 7--9 January, Berkeley,
in liquefiable soils against buckling instability, ISET Journal of USA.
earthquake technology, 43(4), 133--146. IS 1893: part 1. 2002. Criteria for earthquake resistant design of
Bhattacharya, S., Bolton, M.D., and Madabhushi, S.P.G. 2005. A structures: General provisions and structures. 5th Rev. Bureau
reconsideration of the safety of piled bridge foundations in lique- of Indian Standards.
fiable soils. Soils and Foundations, Japanese Geotechnical IS 2911: part 1/sec 2. 1979. Code of practice for design and construc-
Society, 45(4), 13--25. tion of pile foundations. 1st Rev. Bureau of Indian Standards
Bhattacharya, S., Madabhushi, S.P.G., and Bolton, M.D. 2004. An (Reaffirmed 1997).
alternative mechanism of pile failure in liquefiable deposits during Ishihara, K. 1993. Rankine Lecture: Liquefaction and flow failure
earthquakes. Geotechnique, 54(3), 203--213. during earthquakes, Geotechnique, 43, No-3, pp 351--415.
Boulanger, R.W., and Tokimatsu, K., eds. 2005. Proceedings of the Ishihara, K. 1997. Terzaghi oration: Geotechnical aspects of the 1995
workshop on Simulation and Seismic Performance of Pile Kobe earthquake. Proceedings of ICSMFE. Hamburg: 2047--
Foundations in Liquefied and Laterally Spreading Ground, GSP 2073.
145. Davis, CA: ASCE. Itasca Consulting Group. 2006. FLAC, fast Lagrangian analysis of
Byrne, P. 1991. A cyclic shear-volume coupling and pore-pressure continua. version 5.0. Minneapolis, MN: Itasca Consulting Group.
model for sand. In Proceedings of Second International JRA. 1996. Specification for Highway Bridges. Part V, Seismic
Conference on Recent Advances in Geotechnical Earthquake Design. Japanese Road Association.
Engineering and Soil Dynamics (St. Louis, Missouri, March), Kaynia, A.M., and Kausel, K. 1991. Dynamics of piles and pile groups
Paper No. 1.24, 47--55. in layered soil media. Soil Dynamics and Earthquake
Davisson, M.T., and Robinson, K.E. 1965. Bending and Buckling of Engineering, 10(8), 386--401.
Partially Embedded Piles. In Proc. 6th International Conference Kudo, K., Uetake, T., and Kanno, T. 2000. Re-evaluation of nonlinear
on Soil Mechanics and Foundation Engineering, Canada, Vol. III, site response during the 1964 Niigata earthquake using the strong
Div. 3--6, 243--246. motion records at Kawagishi-cho, Niigata City. Proc. 12WCEE,
Donovan, K., Pariseau, W.G., and Ceepak, M. 1984. Finite element Auckland, New Zealand, Paper No. 0969.
approach to cable bolting in steeply dipping VCR slopes. In Lin, S.S., Yu-Ju Tseng, Chen-Chia Chiang and Hung, C.L. 2005.
Geomechanics Application in Underground Hardrock Mining. Damage of piles caused by lateral spreading -- back study of
New York: Society of Mining Engineers, 65--90. three cases. In Boulanger and Tokimatsu, eds. Proceedings of
Eurocode 8. 1998. Design provisions for earthquake resistance of the workshop on Simulation and Seismic Performance of Pile
structures. Part 5: Foundations, retaining structures and geotech- Foundations in Liquefied and Laterally Spreading Ground.
nical aspects. Brussels: European Committee for Standardization. ASCE Geotechnical Special Publication 145.
Finn, W.D. 2005. A study of piles during earthquakes: Issues of design Liyanapathirana, D.S., and Poulos, H.G. 2005. Seismic lateral
and analysis. Bulletin of Earthquake Engineering, 3, 141--234. response of piles in liquefying soil. Journal of Geotechnical and
Finn, W.D.L., and Fujita, N. 2002. Piles in liquefiable soils: seismic Geoenvironmental Engg., ASCE, 131(12) 1466--1479.
analysis and design issues. Soil Dynamics and Earthquake Martin, G.R., Finn, W.D.L., and Seed, H.B. 1975. Fundamentals of
Engineering, 22, 731--742. liquefaction under cyclic loading. J. Geotech., Div. ASCE,
Fukuoka, M. 1966. Damage to civil engineering structures. Soils and 101(GT5), 423--438.
Foundations, Tokyo, Japan, 6(2), 45--52. Meyerhof, G.G. 1957. Discussion, Proceedings of the 4th ICSMGE,
Goh, S., and O’Rourke, T.D. 1999. Limit state model for soil--pile 3, 110.
interaction during lateral spread. In Proc. 7th US--Japan Miwa, S., Ikeda, T., and Sato, T. 2006. Damage process of pile
Workshop on Earthquake Resistant Design of Lifeline foundation in liquefied ground during strong ground motion. Soil
Facilities and Countermeasures against Soil Liquefaction. Dynamics and Earthquake Engg., 26, 325--336.
Seattle, Technical Report MCEER-99-0019, Multidisciplinary National Earthquake Hazards Reduction Program, NEHRP. 2000.
Center for Earthquake Engineering Research, University of Commentary Federal FEMA 369 (USA) on seismic regulations
Buffalo (O’Rourke, Bardet and Hamada, eds): 237--260. for new buildings and other structures.
Hamada, M. 1992a. Large ground deformations and their effects on Rankine, W.J.M. 1866. Useful rules and tables. London: C. Griffin
lifelines: 1964 Niigata earthquake. Case studies of liquefaction and Co.
and lifelines performance during past earthquake. Technical Seed, H.B., and Idriss, I.M. 1971. Simplified procedure for evaluating
Report NCEER-92--0001, Vol. 1, Japanese case studies. soil liquefaction potential. Journal of Soil Mechanics and
Buffalo, NY: National Centre for Earthquake Engineering Foundation Division, ASCE, 97, Proc. Paper 8371, 1249--1273.
Research. Shanker, K., Basudhar, P.K., and Patra, N.R. 2007. Buckling of piles
Hamada, M. 1992b. Large ground deformations and their effects on under liquefied soil conditions. Geotechnical and Geological
lifelines: 1983 Nihonkai-Chubu earthquake. Case studies of lique- Engineering, 25, 303--313.
faction and lifelines performance during past earthquake. Technical Tazoh, T. 2007. Earthquake engineering research on pile founda-
Report NCEER-92--0001, Vol. 1, Japanese case studies. Buffalo, tions with emphasis on pile foundations subjected to large
NY: National Centre for Earthquake Engineering Research. ground deformations. In Bhattacharya, ed. Design of founda-
Hamada, M. 2000. Performances of foundations against liquefaction- tions in seismic areas: Principles and Applications. NICEE,
induced permanent ground displacements, Paper 1754, Proc. of 228--254.
the 12th World Conference on earthquake engineering. Auckland, Timoshenko, S.P., and Gere, J.M. 1961. Theory of elastic stability.
New Zealand. New York: McGraw-Hill.
Geomechanics and Geoengineering 143

Tokimatsu, K., and Suzuki, H. 2004. Pore water pressure response speciality conference, Geotechnical Earthquake Engineering and
Downloaded By: [Sivakumar Babu, G.L.] At: 07:07 24 May 2008

around pile and its effects on p-y behavior during soil liquefaction. Soil Dynamics III, ASCE Geotechnical Special
Soils and Foundations. Japanese Geotechnical Society, 44(6), Publication 75. Reston, VA: American Society of Civil Engineers,
101--110. 1175--1186.
Tokimatsu, K., Mizuno, H., and Kakurai, M. 1996. Building damage Wu, G., and Finn, W.D. 1997a. Dynamic elastic analysis of pile
associated with geotechnical problems. Soils and Foundations, foundations using finite element method in the frequency domain.
Japanese Geotechnical Society, special issue (Jan), 219--234. Canadian Geotech. J., 34, 34--43.
Tokimatsu, K., Hiroshi, O.-O., Satake, K., Shamoto, Y., and Asaka, Y. Wu, G., and Finn, W.D. 1997b. Dynamic nonlinear analysis of pile
1998. Effects of lateral ground movements on failure patterns of foundations using finite element method in the time domain.
piles in the 1995 Hyogoken-Nambu earthquake. Proceedings of a Canadian Geotech. J., 34, 44--52.

You might also like