You are on page 1of 40

Bull Earthquake Eng (2008) 6:407–446

DOI 10.1007/s10518-008-9068-3

ORIGINAL RESEARCH PAPER

A critical review of methods for pile design in seismically


liquefiable soils

S. Bhattacharya · S. P. G. Madabhushi

Received: 3 July 2007 / Accepted: 2 May 2008 / Published online: 10 June 2008
© Springer Science+Business Media B.V. 2008

Abstract Collapse and/or severe damage to pile-supported structures are still observed in
liquefiable soils after most major earthquakes. Poor performance of pile foundations remains
a great concern to the earthquake engineering community. This review paper compares and
contrasts the two plausible theories on pile failure in liquefiable soils. The well established
theory of pile failure is based on a flexural mechanism; where the lateral loads on the pile (due
to inertia and/or lateral spreading) induce bending failure. This theory is well researched in
the recent past and assumes that piles are laterally loaded beams. A more recent theory based
on buckling instability treats the piles as laterally unsupported slender columns in liquefiable
soils and investigates the buckling instability (bifurcation). The objective of this paper is
to investigate the implications to practical pile foundation design that flow from both these
theories. Provisions for design made by major international codes of practice for pile design
including the Japanese Highway Code (JRA) will be considered. The necessity for such codes
to consider alternative forms of failure mechanisms such as the buckling instability of piles
in liquefied ground will be discussed.

Keywords Lateral spreading · Buckling instability · Bending mechanism · Plastic hinge ·


Showa Bridge · Codes of practice

S. Bhattacharya–Previously Departmental Lecturer in Engineering Science, University of Oxford, UK and


Fellow of Somerville College, Oxford. S. P. G. Madabhushi–Fellow of Girton College, Cambridge.

S. Bhattacharya
Dynamics, University of Bristol, Bristol, UK
S. Bhattacharya (B)
Department of Civil Engineering, University of Bristol, University Walk, Bristol, BS8 1TR, UK
e-mail: S.Bhattacharya@bristol.ac.uk; Subhamoy.Bhattacharya@eng.ox.ac.uk
S. P. G. Madabhushi
Geotechnical Engineering, University of Cambridge, Cambridge, UK

123
408 Bull Earthquake Eng (2008) 6:407–446

1 Introduction

Buildings and bridges on loose to medium dense sands are often built on piles to limit
settlements because the surface ground layers are often not stiff enough to support the struc-
tures. In an earthquake if these loose sands are saturated, they lose strength as excess pore
water pressure is generated and the soil tends to liquefy. This means that if the soil is on a
slope it will flow downslope which is often termed as “lateral spreading”. Lateral spreading
is a term used to represent the permanent lateral ground displacement after an earthquake.
It is often found after an earthquake that the buildings supported on piles have collapsed or
suffered excessive damage; see for example Photographs 1 and 2.
The building at Kandla Port in Photograph 1 was located in a laterally spreading ground
and it suffered significant rotation following the 2001 Bhuj earthquake. This building was
supported on pile foundations that extended into stiff clay with very little end bearing,
Madabhushi et al. (2005). Evidence of lateral spreading of ground was observed in the
close proximity to this building. The building in Photograph 2 was founded in level ground
and no lateral spreading was observed at the site. It is generally acknowledged that the failure
of these buildings was caused by the lateral earth pressure applied by the laterally spreading
liquefied sand and any non-liquefied stabilised crust resting on the top of the liquefied soil.
As a consequence much of the research in this area was focussed on;
1. Establishing lateral spreading as a mechanism of failure for pile foundations
2. Determining methods for estimating the lateral loads attracted by the pile foundations
so that geotechnical engineers can design for those lateral loads.
This mode of research has lead to a significant body of research, for example,
Abdoun and Dobry (2002), Hamada (1992a,b, 2000), Tokimatsu et al. (1996, 1997, 1998),
Ishihara (1997), Finn and Thavaraj (2001), JRA (2002), Finn and Fujita (2002), Tazoh (2007).
The down-slope deformation of the ground surface adjacent to the pile foundation seems to

FLOWING SOIL

Water Table Non-liquefied


stabilised crust

Liquefied
Depth of layer
liquefaction

Fig. 1 Current understanding of pile failure

123
Bull Earthquake Eng (2008) 6:407–446 409

support this explanation. Figure 1 explains the hypothesis of failure. This mechanism is
therefore based on flexure of the pile causing a bending failure of the pile foundation. This
mechanism forms the basis for the design guidelines developed by JRA (1996, 2002). Dobry
et al. (2003) and Abdoun et al. (2003) recently suggested that in the presence of non-liquefied
crust, the lateral loading created by liquefied ground can be ignored as the lateral passive
earth pressures created in the non-liquefied crust dominate the lateral loading generated on
the pile foundation.
However, recent research by the authors has suggested an alternative explanation
recognising the fact that pile foundations are normally carrying significant axial loads at
the time of the earthquake. When the soil around the pile liquefies it loses much of its stiff-
ness and strength, so the piles now act as unsupported long slender columns, and simply
buckles under the action of the axial (superstructure) loads. The stress in the pile section
will initially be within the elastic range, and the buckling length will be the entire ‘unsup-
ported’ length in the liquefied soil. Lateral loading, due to slope movement, inertia or any
out-of-alignment eccentricities will increase lateral deflections. When the pile suffers suffi-
cient lateral deflections plastic hinges can form, reducing the buckling load, and promoting
more rapid collapse. Therefore, this hypothesis is based on a buckling mechanism.
In order to establish which of the above two failure mechanisms are causing pile
foundation failures, it is necessary to revisit some of the observed pile foundation failures in
recent earthquakes. The next section describes the observed damage to the piles following
excavation of some of the damaged piled foundations from the literature.

Photograph 1: Kandla Port tower after the 2001 Bhuj earthquake. The building is located in a laterally spreading
ground; Photograph 2: Collapse of a building in level ground following the 1995 Kobe earthquake

2 Observed pile damage from detailed field investigations

As earthquakes are very rapid events and as much of the damage to piles occurs beneath
the ground, it is hard to ascertain the failure mechanism unless deep a excavation is car-
ried out. Twenty years after the 1964 Niigata earthquake and also following the 1995 Kobe
earthquake, investigations has been carried out to find the failure pattern of the piles, see
Yoshida and Hamada (1990), BTL Committee (2000), Tokimatsu and Asaka (1998), Tazoh
(2007). Piles were excavated or extracted from the subsoil, borehole cameras were used to

123
410 Bull Earthquake Eng (2008) 6:407–446

take photographs, and pile integrity tests were carried out. These studies hinted the location
of the cracks and damage patterns for the piles. Of particular interest is the formation of
plastic hinges in the piles. This indicates that the stresses in the pile during the earthquakes
exceeded the yield stress of the material of the pile. As a result, design of pile foundation
in seismically liquefiable areas still remains a constant source of attention to the earthquake
geotechnical engineering community.
The main observations are:
1. The failure of foundations not only occurred in laterally spreading grounds but also
was observed in level grounds where no lateral spreading would be anticipated; see
Photographs 1 and 2. As mentioned earlier, the building in Photograph 1 was located
in laterally spreading ground and the building in Photograph 2 was located in a level
ground.
2. The failure patterns of the buildings in level ground and in laterally spreading ground are
similar i.e. the buildings tilt considerably, see for example Photograph 1 and 2. In most
cases, the superstructure remains undamaged. Therefore, by considering the building
failure alone, it is difficult to ascertain the type of ground i.e., whether lateral spreading
ground or level ground.
3. Figure 2a, b, c and d shows several examples of pile damage along which the soil profiles
and the damage pattern. Cracks were observed near the bottom and top boundaries
between liquefied and non-liquefied layers. Often cracks were observed at the pile
head.
4. Plastic hinges also formed not only at the boundaries of the liquefiable and non-liquefiable
layers but also at various depths.
5. Horizontal displacement of the piles were measured in some cases of pile failure, see for
example Yoshida and Hamada (1990), Hamada and O’Rourke (1992). In some cases,
the displacement of the piles was found to agree with the horizontal displacement of the
ground.
6. There are few cases where plastic hinges formed at the middle of the liquefied layer, see
for example Fig. 2b. It has been reported, Yoshida et al. (2005) that this damage was
observed when the structure was very close to the quay wall which moved towards the
sea significantly.
7. It has been observed that large diameter piles performed well during the earthquake. For
example, the older jetties of the Kandla port which were supported on 0.5 m diameter
RCC piles suffered cracking during the 2001 Bhuj earthquake. In contrast, in the same
area, the 1.0 m diameter RCC piles on the newer jetties and the concrete-filled-steel
tubular piles performed well.

2.1 Plausible failure mechanisms for pile foundations in liquefiable soils

As explained in the introduction to this paper, there are two plausible failure mechanisms
that can explain pile failure in liquefiable soils:
1. Bending mechanism where the piles are treated as beams. The lateral loads on the piles
are due to inertial effects from the superstructure and kinematic effects due to ground
movement.
2. Buckling instability where the piles are treated as beam-columns i.e., axially loaded
slender columns carrying lateral loads. The piles are treated as unsupported columns in
the liquefiable zones.
The next section describes each of these mechanisms in more detail.

123
Bull Earthquake Eng (2008) 6:407–446 411

a
Depth (m) SPT N-value
Depth (m)
0 0 0 10 20 30 40 50
GL±0

1 1

2 2 5

3 3
Sand

Depth(m)
4 4 10
5 5

6 6 15 No.1
No.2
7 7

8 8
20

SPT N-value
0 10 20 30
b GL- m
1,700

1 7
2 9
2,500∼3,500

3 8
4 7
5 13
6 7
4,500∼6,500

7 9
≈11,000

φ≈10∼15°
8 10
9 12
10 8
2,000∼3,000

11 19
12 23
400

13 19
14 25
1,000∼2,000
Unit: mm

No. 2 No.3 Depth


c No. 1
Pile cap Pile cap (m) Soil type SPT N-value
0 Pile cap
0 50
0
-5
Reclaimed
-5 fill
-7.30 m -8.50∼
end of -8.90 m -9.17∼-9.39 m
investigation Large cracks large cracks with -10
-10 with concrete concrete failure
Depth (m)

failure
-13.00 m
Silt
pile joint -13.00 m -15 clay
pile joint
-15
-20

-20 -20.80 m
-25
end of
investigation -22.90 m Sand
end of -30 and
investigation
-25 clay
-35

-30 NSN NSN NSN -40

Fig. 2 (a) Building in Niigata city (Yoshida and Hamada 1990). (b) Building in Niigata city (Kawamura
et al. 1985). (c) Building in Kobe city (Fujii et al. 1998). (d) A Governmental building, Kobe city (Editorial
Committee for the Report on the Hanshin-Awaji Earthquake Disaster 1998)

123
412 Bull Earthquake Eng (2008) 6:407–446

Fig. 2 continued d SPT N-value


0 10 20 30 40 50 60
0
0.85
1
7

3.17
1.85
2
11
2.85

Depth (m)
3 B1F
3

4
16

2.40
4.85 pit
5
5.65 Base9
6
9 Crack
7
11 investigation hole
7.75
8
8.65 6 φ1,100
9
3.6
9.70
10 10.15
20

11
38 Crack
11.60
12 12.20
68
12.80
13
15
13.85
14
41
14.95
15
54

16
69

17
45
17.80
18 18.25
48
18.90
19
94
19.85
20 20.20
40
20.75
21 21.15
21.65 13

22
22.55 20

23
23.60 300

3 Theory of pile failure based on bending mechanism

According to the authors’ knowledge, “lateral spreading” was first proposed as a


possible cause of pile failure during liquefaction in a report published by the National
Research Council (1985). This report also identifies that “lateral spreading” is responsi-
ble for more damage during earthquakes than any other form of liquefaction-induced ground
failure. The bending based failure mechanism has been accepted as the explanation of pile
failure in many earthquakes. Hamada (2000) in the 12th World Conference on Earthquake
Engineering concludes that permanent displacement of the non-liquefied soil overlying the
liquefied soil is a governing factor for pile damage. A similar conclusion was also reached
by Berrill et al. (2001). This section of the paper describes the pile failure theory based on
lateral spreading as explained by Ishihara (1997) and Tokimatsu et al. (1998).

123
Bull Earthquake Eng (2008) 6:407–446 413

3.1 Failure theory based on Tokimatsu et al. (1998)

Tokimatsu et al. (1998) schematically described the soil-pile structure interaction in


liquefiable soil as shown in Fig. 3a. The assumptions are:
1. Prior to the development of pore water pressure, the inertia force from the superstructure
may dominate. This is referred to as stage I in Fig. 3.
2. Kinematic forces from the liquefied soil start acting with increasing pore pressure. This
is referred to as stage II in Fig. 3.
3. Towards the end of shaking, kinematic forces would dominate and have a significant
effect on pile performance particularly when permanent displacements occur in laterally
spreading soil.

3.2 Ishihara’s (1997) concept of pile failure (Top-down and bottom-up effect)

Ishihara (1997) in his Terzaghi oration at ICSMGE, Hamburg (International Conference on


Soil Mechanics and Geotechnical Engineering) summarised the seismically induced loading
on the pile by introducing the concepts of ‘top down effect’ and ‘bottom up effect’. These are
described below:

3.2.1 Top-down effect

At the onset of shaking, the inertia forces of superstructure are transferred to the top of the
pile and ultimately to the soil. He assumes that during the main shaking, sandy soils in a
deposit have not softened significantly due to liquefaction and that the relative movement
between the piles and ground are small. However, he postulates that if ground motion is suf-
ficiently high such that the induced bending moment in the piles exceeds the limiting value,
the piles may fail. Since the horizontal load comes from the inertia force of superstructure,
it is referred to as ‘top down effect’. He concludes that the observed failure of a pile in the
upper portion after an earthquake may be attributed to this effect.
Ishihara (1997) also reports: “It has been known that onset of liquefaction takes place
approximately at the same time as the instant when peak acceleration occurs in the course of
seismic load application having an irregular time history”.

3.2.2 Bottom-up effect

In sloping grounds, the softened ground will start to move horizontally following the onset of
liquefaction. Under this condition, lateral forces would be applied to the pile body embedded
in the ground, leading to deformation of the pile in the direction of the slope. He assumes
that seismic motion has already passed the peak but the shaking may still be persistent with
lesser intensity and therefore the inertia force transmitted from the superstructure will be
small. Under such a loading condition, the maximum bending moment induced by the pile
may not occur near the pile head but at a lower portion at some depth and this is referred to
as, ‘bottom-up effect’.

3.3 Japanese highway code of practice (JRA 1996 and 2002 edition)

The Japanese Highway code of practice JRA (1996) has incorporated the concept as of
“TOP-DOWN” and “BOTTOM-UP” effects as shown in Fig. 3a. The code advises practising

123
414 Bull Earthquake Eng (2008) 6:407–446

Fig. 3 (a) Schematic diagram


showing the pile failure (after
Tokimatsu et al. 1998). (b) JRA
(1996) code of practice showing
the idealisation for seismic
design of bridge foundation. (c)
Idealization of the applied force.
(d) Relation between lateral force
and displacement at pile top. (e)
Schematic diagram of the Fall-off
of the girders in Showa Bridge,
Takata et al. (1965). (f) Failed
pile and the soil profile, Fukuoka
(1966). (g) Schematic diagram
showing the predicted loading
based on JRA code

123
Bull Earthquake Eng (2008) 6:407–446 415

f g
Water level 9m
Mud line
3m (assumed)

9kPa (based A
on JRA1996)

69kPa (based 10m


on JRA1996)

Dense non- 6m
liquefiable
soil

Fig. 3 continued

engineers to design piles against bending failure assuming that the non-liquefied crust exerts
passive earth pressure on the pile and the liquefied soil offers 30% of total overburden pressure.
Many researchers, as will be discussed later on have verified their experimental results against
such pressure distribution. The code also advises designers to check against bending failure
due to kinematic forces and inertia separately, i.e., a check against bending failure due to the
combination of the two loads (Inertial and Kinematic) is not encouraged. The basis of such
clauses can be well understood from the “Ishihara’s top down and bottom up concepts”.

3.4 Basis of formulation of the design pressure laid down in the JRA (1996) code

Yokoyama et al. (1997) reports that the JRA (1996 or the 2002 edition) code was
formulated by back-analysing a few piled bridge foundations of the Hanshin expressway that

123
416 Bull Earthquake Eng (2008) 6:407–446

were not seriously damaged following the 1995 Kobe earthquake. Specifically, the bridge
foundation suffered residual horizontal displacements. One such bridge foundation is the
Pier P-216 which was located at the north edge of Rokko Island. Some details of the analysis
are presented in this section:

3.5 Case study of the performance of the P-216 pier during the 1995 Kobe earthquake

The pier P-216 was a 2-story steel rigid frame pier and was supported by 1.5 m diameters
cast-in-place concrete piles. There were two bearings at the top of the pier-cap supporting the
girders of the bridge. One of the bearing was movable and the other was fixed. The soil at the
site consists of sandy alluvial fill, alluvial clay and alternation of sand and clay. Following
the earthquake, the residual horizontal displacement of this pier was 0.9 m i.e., this can be
viewed as the apparent final pile head displacement. The ground water table at the site was
3.3 m below the ground level and analysis confirmed that liquefaction occurred in the sandy
artificial fill below the ground level.
Forces that could cause 0.9 m of residual displacement were estimated based on a Push-
Over type analysis and is shown in Fig. 3c and d. Figure 3c describes the idealization of the
applied force and foundation of the P-219. Figure 3d graphically shows the relationship of
the lateral force required for a particular amount of pile head displacement.
The assumptions in the analysis are:

1. The layer near the ground surface which is 3.3 m deep did not liquefy but did move with
the liquefied layer. The liquefied layer translated downslope.
2. Both the non-liquefied crust and the liquefied soil exerted force on the bridge foundation
causing the residual displacement of 0.9 m.
3. The force applied by the non-liquefied crust was assumed to be passive and the unknown
quantity is the force exerted by the liquefied soil.
4. The force applied by the liquefied soil has been varied until displacement of 0.9 m
could be matched. Based on trial and error, it was found that a ratio of the force
applied in the liquefied layer to the overburden pressure 0.32 could predict the pile head
displacement.

Similar analysis was conducted for four other bridge piers on the Route 5 of the Hanshin
Expressway, and a factor of 0.3 could predict the residual pile head deflection. This led to
the pressure distribution in JRA (1996) code, see Fig. 3b.

3.6 A particular case study: Showa Bridge failure during the 1964 Niigata earthquake

This section describes the Showa Bridge and the damage that occurred during the 1964
Niigata earthquake, (Fukuoka 1966), see Photograph 3. It will be shown that the piles sat-
isfy the calculation criteria of the JRA (1996) code, i.e., have enough apparent strength to
resist lateral spreading, but they failed. The bridge was built over the Shinano River and was
completed just a month before the earthquake.
The bridge had a width of 24 m and total length of 303.9 m. The superstructure of the
bridge consisted of 12 composite girders. The foundations of each pier consisted of a single
row of 9 tubular steel piles connected laterally as shown in Photograph 1. After the earth-
quake five girders (G3 to G7) fell into the river as shown in Fig. 3e. Figure 3f shows the
post earthquake failure investigation and recovery of a damaged pile, together with the soil
investigation data. Table 1 summarises the design data of the pile.

123
Bull Earthquake Eng (2008) 6:407–446 417

Table 1 Design data of pile


Length 25 m
External diameter 609 mm
Internal diameter 591 mm
Material Steel
E (Young’s Modulus) 210 GPa

Photograph 3: Collapse of Showa Bridge [Photo courtsey: NISEE]

3.7 Eyewitness report

According to reliable eyewitnesses, “the girders began to fall somewhat later, perhaps about 0
to 1 min after the earthquake motion ceased”, Hamada (1992a,b). As can be seen from Fig. 3e,
the piles of pier P5 deformed towards the left and the piles of pier P6 deformed towards the
right (Fukuoka 1966; Takata et al. 1965). Recently, Yoshida et al. (2007) have revisited the
failure of the Showa Bridge and collated many more eyewitness accounts from the survivors
of the Niigata earthquake of 1964 and confirm the post earthquake failure of the bridge.

3.8 Why did pier P5 deform in the direction away from the slope?

Had the cause of pile failure been due to lateral spreading, as suggested by Hamada (1992a,b)
the piers should have displaced identically in the direction of the slope. Pier P5 would have
deformed in towards the middle of the river. Furthermore, the piers close to the riverbanks
or the abutments piles did not fail, even though the lateral spread was seen to be severe at
this location. This creates doubt in the explanation of the bridge failure by lateral spreading.

3.9 Bending calculation based on JRA (1996 or 2002) code

Photograph 3 and Fig. 3f indicate that the soil surrounding the failed piles was fully submerged
sand, so it would be unwarranted to invoke a non-liquefied crust at the top of the liquefied soil.
Figure 3g shows the design lateral loading diagram based on JRA (1996) code i.e., Fig. 3b.
The calculation below estimates the maximum bending moment ignoring axial load effects.
Assume the bulk unit weight of soil is 20 kN/m3 .
Maximum lateral spreading pressure at mudline at point A in Fig. 3g
= 30% of total overburden pressure due to water = 0.3 × 10 kN/m3 × 3 m
= 9 kPa.
Maximum lateral spreading pressure at 10 m depth acting at point B in Fig. 3g
= 30% of total overburden pressure = 0.3 × (20 kN/m3 × 10 m+10 kN/m3 ×3 m)
= 69 kPa.

123
418 Bull Earthquake Eng (2008) 6:407–446

Maximum bending moment, at point B in Fig. 3g, due to spreading forces (trapezoidal
loading)
= (0.5 × 60 kPa × 10 m × 0.609 m × 3.33 m)+(9 kPa × 10 m × 0.609 m × 5 m)
= 608 kNm + 274 kNm
= 882 kNm.
The plastic moment capacity of the section AB (9 mm thick)
 
0.6093 0.5913
− m3 × 500 MPa = 1620 kNm
6 6
Hence the calculated factor of safety against plastic bending failure, ignoring axial load
effects, is (1620/882) = 1.84. Thus, according to this provision of JRA code, the bridge
should not have collapsed.

4 Limitations of the theory of pile failure based on bending mechanism

1. The basic assumption of pile failure is obtained by treating the pile as a beam element and
assuming that the lateral loads due to inertia and slope movement cause bending failure
in the pile. The effect of axial load in promulgating such a failure due to the horizontal
displacement induced by lateral spreading or inertial loading has been ignored.
2. It has been argued that as the horizontal displacement of the ground in some cases is
found to agree with displacement of the pile head, the damage to the piles is therefore
linked to the liquefaction induced ground flow (See the section on observed pile damage
and also Yoshida et al. (2005)). This observation has led to the hypothesis that lateral
ground movement may have caused the pile failure. Equally, the lateral spreading ground
can induce horizontal displacement at the pile head thereby creating large moment loads
due to P-δ effect. This, together with loss of lateral stiffness of the soil can cause the pile
to fail by buckling instability.
3. The analysis of the Showa Bridge seems to suggest that according to JRA (1996) based
calculations indicate that the Showa Bridge should have survived the Niigata earthquake,
while buckling instability based calculations indicate that the Showa Bridge should have
collapsed (as will be shown in the section below). The fact that the Showa Bridge actually
did collapse seem to offer some credence to the buckling based argument.

5 Theory of pile failure based on buckling instability

Bhattacharya (2003) and Bhattacharya et al. (2004) proposed an alternative mechanism of


pile failure. This failure mechanism, based on buckling instability theory, has been formu-
lated by back-analysing 15 cases of pile foundation performance and verified by high quality
experiments (Dynamic Centrifuge Tests). The next section explains the theory.
During earthquakes, soil layers overlying the bedrock are subjected to seismic excitation
consisting of numerous incident waves, namely dilatational shear (S) waves, pressure (P)
waves, and surface (Rayleigh and Love) waves which result in ground motion. The ground
motion at a site will depend on the stiffness characteristics of the layers of soil overlying
the bedrock. This motion will also affect a piled structure. As the seismic waves arrive in
the soil surrounding the pile, the soil layers will tend to deform. This seismically deforming
soil will try to move the piles and the embedded pile-cap with it. Subsequently, depending

123
Bull Earthquake Eng (2008) 6:407–446 419

upon the rigidity of the superstructure and the pile-cap, the superstructure may also move
with the foundation. The pile may thus experience two distinct phases of initial soil-structure
interaction.
1. Before the superstructure starts oscillating, the piles may be forced to follow the soil
motion, depending on the flexural rigidity (EI) of the pile. Here the soil and pile may
take part in kinematic interplay and the motion of the pile may differ substantially from
the free field motion. This may induce bending moments in the pile.
2. As the superstructure starts to oscillate, inertial forces are generated. These inertia forces
are transferred as lateral forces and overturning moments to the pile via the pile-cap. The
pile-cap transfers the moments as varying axial loads and bending moments in the piles.
Thus the piles may experience additional axial and lateral loads, which cause additional
bending moments in the pile.
These two effects occur with only a small time lag. If the section of the pile is inadequate,
bending failure may occur in the pile. The behaviour of the pile at this stage may be approxi-
mately described as a beam on an elastic foundation, where the soil provides sufficient lateral
restraint. The available confining pressure around the pile is not expected to decrease sub-
stantially in these initial phases. The response to changes in axial load in the pile would not
be severe either, as shaft resistance continues to act. This is shown in Fig. 4a (Stage II).
1. In loose saturated sandy soil, as the shaking continues, pore pressure will build up and
the soil will start to liquefy. With the onset of liquefaction, an end-bearing pile passing
through liquefiable soil will experience distinct changes in its stress state.
2. The pile will start to lose its shaft resistance in the liquefied layer and shed axial loads
downwards to mobilise additional base resistance. If the base capacity is exceeded, set-
tlement failure will occur.
The liquefied soil will begin to lose its stiffness so that the pile acts as an unsupported column
as shown in Fig. 4a (Stage III). Piles that have a high slenderness ratio will then be prone to
axial instability, and buckling failure will occur in the pile, enhanced by the actions of lateral
disturbing forces and also by the deterioration of bending stiffness due to the onset of plastic
yielding. Dynamic centrifuge tests, study of case histories and analytical work carried out
by Bhattacharya (2003), Bhattacharya et al. (2004) has conclusively shown the above failure
mechanism. This particular mechanism is currently missing in all codes of practice.
In sloping ground, even if the pile survives the above load conditions (i.e. safely carry
the axial load and the lateral inertial loads at fully-liquefied condition), it may experience
additional drag load due to the lateral spreading of soil. Under these conditions, the pile may
behave as a beam-column (column with lateral loads); see Fig. 4a (Stage IV). This bending
mechanism is currently considered most critical for pile design, see for example JRA (1996
or 2002).

5.1 Study of case histories to verify the theory of pile failure by buckling instability

In this study, fifteen reported cases of pile foundation performance during earthquake-
induced liquefaction have been studied and analysed. Six of the piled foundations were
found to survive while the others suffered severe damage (Bhattacharya et al. 2004). The
analysis is carried out based on the simple principle that “A pile is an unsupported slen-
der column in the liquefiable zone”. The parameters rmin (minimum radius of gyration)
and Leff (effective buckling length of the pile) are introduced to analyse the piles,
as follows.

123
420 Bull Earthquake Eng (2008) 6:407–446

a Pstatic Pstatic+ Pdynamic Pstatic+ Pdynamic Pstatic+ Pdynamic

Plateral Plateral Plateral

Non Liquefied
Crust

Loose Sand Liquefied


Sand

Dense
Sand

Stage-I: Stage-II: Stage-III: Stage-IV:


Before earthquake in a Shaking starts. Soil Soil liquefied. Inertia In sloping ground
level ground yet to liquefy. Pile forces may act. Pile acts as lateral spreading may
acts as a beam a column and may buckle start
or settle down

b Pile head – fixed but pile


pinned at the base

Pile head – free to


Pile head – translate but restrained
unrestrained against rotation

Leff = 2Lo

Leff = Lo

Leff = 2Lo

c
Structure on pile
foundations

Liquefiable Leff = Lo
soil

Dense soil

Fig. 4 (a) Different stages of the loading. (b) Concept of effective length. (c) Effective length of pile in case
of a large raft supported on piles. (d) Effective length of piled foundation likely to be encountered in practice.
(e) Leff versus rmin for piles studied, Bhattacharya et al. (2004). (f) Plot of P and Pcr of the poor performance
piles in Table 3 and Fig. 4(e). (g) Amplification factors of piles for lateral loads. (h) Plot of concrete pile
performance mentioned in Table 3. (i) Schematic of the Showa Bridge deck (adapted from Iwasaki 1984). (j)
Limiting axial loads for the Showa Bridge Piles—Plastic squash load PP , Euler’s elastic buckling load Pcr ,
Rankine’s elastic–plastic buckling load PF (k) Observed stiffness of liquefied soil in an element test, Yasuda
et al. (1999)

123
Bull Earthquake Eng (2008) 6:407–446 421

d
Non-liquefied Non-liquefied
crust crust

Non-liquefied
crust

Liquefiable Liquefiable
soil (L0 )
soil (L0 )

Dense Dense Dense


Case 1 Case 2 Case 3

Liquefiable
Liquefiable Liquefiable soil (L 0)
soil (L0 ) soil (L0 )

Dense Dense
Dense
Ca se 4 Case 5 Case 6

f 3000

2500

2000
Pcr (kN)

1500

1000 Poor
performance
P/Pcr = 1
500
P/Pcr = 0.5

0
0 500 1000 1500 2000
P (kN)

Fig. 4 continued

123
422 Bull Earthquake Eng (2008) 6:407–446

g
Timoshenko and Young (1961)
Halabe and Jain (1993) Fixed-headed semi-infinite and finite piles
Halabe and Jain (1993) Free-free rigid pile
Halabe and Jain (1993) Free headed semi-infinite and finite piles

17
Buckling amplification factor

13

1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(P/Pcr)
h
21
Normal Stress in the pile (MPa)

Euler's Theoretical curve for


M25 concrete
16 Good performance

Poor performance

10 Yield stress line

Rankine's formula

0
0 50 100 150 200 250 300 350
(Leff/rmin)

i
24.8m

200 mm slab thickness, 1.5m


50 mm asphalt Crash
barrier

1.3m

0.7m

Fig. 4 continued

rmin : The minimum radius of gyration of the pile section about any axis of bending (m).
This parameter can represent piles of any shape (square, tubular or solid circular) and is used
by structural engineers for studying buckling instability. It is given by:

1
rmin = , (1)
A
where I is the second moment area of the pile section about the weakest axis (m4 ); A is area
of the pile section (m2 ).
For solid circular piles rmin is 0.25 times the diameter of the pile and for tubular piles rmin
is approximately 0.35 times the outside diameter of the pile.

123
Bull Earthquake Eng (2008) 6:407–446 423

j Limit loads of a 600mm dia tubular steel cantilver


9000
PP
8000

7000
axial load P kN

PF
6000

5000

4000
Pcr
3000
Pcr/2
2000

1000

0
0 5 10 15 20 25 30
unsupported length m

Fig. 4 continued

Leff : The effective buckling length of the pile, taken from column stability theory. Leff is
also familiar as “Euler’s buckling length” of a strut pinned at both ends. The definition of

123
424 Bull Earthquake Eng (2008) 6:407–446

effective length of piles is shown in Fig. 4b. Here, it depends on the depth of the unsupported
region (Lo) and the boundary fixity conditions of the pile base and pile head as shown in
Fig. 4b. The particular case of a piled raft with piles fixed in direction at the top and bottom
of a liquefiable zone, and where Leff = Lo, is shown in Fig. 4c.
Standard charts for estimating the effective length of slender columns can equally be used
to find the effective length of piles. Table 2 lists the different cases for ready reference.
The assumption of pile fixity calls for significant lateral restraint in the underlying ground,
which could be assured only if the piles are grouted into a rock-socket. Fixity at the junction
between liquefiable and non-liquefiable soils is, however, impossible. It must be recalled that
the base of a liquefied soil layer is at zero effective stress, and therefore acts temporarily as
a free surface for any underlying soil. Fixity requires a pile penetration of approximately 5
diameters into dense soil. In practice, designers should make a corresponding allowance for
imperfect fixity when they select a value for the length Lo that could be left unsupported due
to earthquake-induced liquefaction.

5.2 Buckling load of the pile

The Euler’s buckling load of the pile (Pcr ) is calculated from the well known formula given
by Eq. 2.
π2
Pcr = EI, (2)
L2eff
where EI is the Bending stiffness of the pile and Leff is effective length of the pile defined in
the earlier section.
Figure 4d shows some cases of piled foundations likely to be encountered in practice. The
fixity at the bottom of the liquefied soil can be classified into two cases:
1. Embedment more than 5 times the diameter of the pile, such as Cases 1, 3, 4 and 6 in
Fig. 4d. In such case, that the slope of the deflected pile is zero at the bottom of the
liquefied layer.
2. Less embedment and therefore the pile can rotate at the bottom of the liquefied soil.
Examples are Cases 2 and 5 in Fig. 4d.
Apart from Table 2 and Fig. 4d, other boundary conditions of the pile at the top and bottom
of the liquefiable soil may exist and similar expressions can be worked out.
Table 3 summarizes the parameters for 15 case histories of piled foundation performance
based on buckling parameters. The ratio Leff /rmin is termed the slenderness ratio of the pile
in the region that could become unsupported. Figure 4e plots the Leff against the rmin of
the pile sections with identification of their performance during earthquakes. A line rep-
resenting a slenderness ratio (Leff /rmin ) of 50 is drawn and it distinguishes between unac-
ceptable and acceptable pile performance. This line is of some significance in structural
engineering, as it is often used to distinguish between “long” and “short” columns. Columns
having slenderness ratios below 50 are expected to fail by plastic squashing whereas those
above 50 are expected to fail by buckling, both modes being modified by induced bending
moments. This slenderness ratio of 50 signifies length to diameter of about 12 for RCC
columns.
This study essentially implies:
1. Piles that will be considered as short concrete columns performed well irrespective of
the type of ground i.e., be it level ground or lateral spreading ground.
2. Piles that would be considered as long columns collapsed during earthquakes.

123
Table 2 Various cases of effective length of piles in liquefiable soils

Case ID in Boundary condition of the pile at the top and bottom of the Effective length Buckling load of each Example
Fig. 4d liquefied layer pile
Top Bottom

Case 1 Fixed [θ = 0, δ = 0] Fixed [Sufficient embedment Leff = 0.5L0 4π 2 EI Pile groups with raked
L20
at the dense layer] [θ = 0, piles
δ = 0]
Large non-liquefied
crust which will
Bull Earthquake Eng (2008) 6:407–446

not slide
Case 2 Free to translate but restrained Pinned [Insufficient embed- Leff = 2L0 π 2 EI NFCH building,
4L20
against rotation–sway ment at the dense layer] Hamada (1992a,b)
frame [θ = 0, δ  = 0] [θ  = 0, δ = 0]
Case 3 Free to translate but restrained Fixed [Sufficient embedment Leff = L0 π 2 EI Most cases fall under
L20
against rotation–sway at the dense layer] [θ = 0, such category
frame [θ = 0, δ  = 0] δ = 0]
Case 4 Fixed in direction but free to Fixed [Sufficient embedment Leff = 0.7L0 2π 2 EI Pile groups with raked
L20
rotate [θ  = 0, δ = 0] at the dense layer] [θ = 0, piles. Improper
δ = 0] pile-pilecap con-
nection
Case 5 Fixed in direction but free to Pinned [Less embedment at Leff = L0 π 2 EI Pile groups with raked
L20
rotate [θ  = 0, δ = 0] the dense layer] [θ  = 0, piles. Improper
δ = 0] pile-pilecap
connection
Case 6 Free i.e. unrestrained against Fixed [Sufficient embedment Leff = 2L0 π 2 EI Piles in a row such as
4L20
rotation and displacement at the dense layer] [θ = 0, the Showa Bridge
[θ  = 0, δ  = 0] δ = 0] piles
Case 1: In such cases, it is necessary to verify if the non-liquefied crust at the top of the liquefied soil can slide laterally and load the pile. If the non-liquefied crust can slide, the
unsupported length will be (L0 + the thickness of the non-liquefied crust) and will be similar to Case 3

123
425
426 Bull Earthquake Eng (2008) 6:407–446

Table 3 Summary of case histories, Bhattacharya (2003)

ID in Fig. 4e Case history and reference Pile section/type L0 * (m) Leff (m) rmin (m)

A 10 storey-Hokuriku building, 0.4 m dia RCC 5 5 0.1


Hamada (1992a)
B Landing bridge, Berrill et al. 0.4 m square PSC 4 2 0.12
(2001)
C 14 storey building, Tokimatsu 2.5 m dia RCC 12.2 12.2 0.63
et al. (1996)
D Hanshin expressway pier, 1.5 m dia RCC 15 15 0.38
Ishihara (1997)
E LPG tank 101, Ishihara (1997) 1.1 m dia RCC 15 15 0.28
F Kobe Shimim hospital, Soga 0.66 m dia steel tube 6.2 6.2 0.23
(1997)
G N.H.K building, 0.35 m dia RCC 10 20 0.09
Hamada (1992a)
H NFCH building, 0.35 m dia RCC hollow 8 16 0.10
Hamada (1992a)
I Yachiyo Bridge 0.3 m dia RCC 8 16 0.08
Hamada (1992a)
J Gaiko Ware House, Hamada 0.6 m dia PSC hollow 14 28 0.16
(1992b)
K 4 storey fire house, Tokimatsu 0.4 m dia PSC 18 18 0.10
et al. (1996)
L 3 storied building at Kobe 0.4 m dia PSC 16 16 0.12
university, Tokimatsu et al.
(1998)
M Elevated port liner railway, 0.6 m dia RCC 12 12 0.15
Soga (1997)
N LPG tank–106,107 0.3 m dia RCC hollow 15 15 0.08
Ishihara (1997)
O Showa Bridge, 0.6 m dia steel tube 19 38 0.21
Hamada (1992a)
L0 * = Length of pile in liquefiable layer/buckling zone, see Fig. 4b, Table 2
PSC: Pre-Stressed Concrete; RCC: Reinforced Cement Concrete

5.3 Comment on the quantitative study of case histories

In general, those reporting pile failures [References in column 2 of Table 3] have focussed on
lateral loading issues, rather than axial load effects (buckling instability). In the cases listed
in Table 3, the average pile axial loads P at the time of the earthquake have been estimated
rather crudely by assuming them equal to the axial loads that could be considered allowable
in conventional designs based on their geometry, and the ground profile at their location.
This is purely from the geotechnical considerations. They are listed in Bhattacharya (2003)
or Bhattacharya et al. (2004).
The plot of P and Pcr (the elastic buckling load, if unsupported) for the piles that failed in
Fig. 4e can be seen in Fig. 4f following Bhattacharya et al. (2004). It may be observed that
the piles that failed had a (P/Pcr ) ratio between 0.5 and 1, and a slenderness ratio (based on
the assumed unsupported length) greater than 70. On the other hand, the analysis of the case
histories shows that piles with a (P/Pcr ) ratio below 0.1 and a slenderness ratio smaller than
50 survived earthquakes even though they carried loads through soils that were spreading
laterally.

123
Bull Earthquake Eng (2008) 6:407–446 427

5.4 Does this P/Pcr ratio of 0.5 signify something?

A pile should not be close to its critical load at full liquefaction. But how close to critical load
can be allowed? In other words, what Factor of Safety against buckling should be allowed?
Pcr are the critical loads at which a structure becomes laterally unstable. Each of these loads
has an associated buckled mode shape which is the shape that the structure assumes in a buck-
led condition, which can be seen in Fig. 4d. The theoretical buckling load can be computed
for the given boundary conditions such as Table 2. However, real-life structures have imper-
fection and the non-linearity in response prevents most structures, if not all, from reaching
their predicted buckling load. In other words, formulas in Table 2 over-predict the buckling
load. The next section explains the important issues related to imperfections in buckling
calculations and the importance of P/Pcr ratio.

5.5 Amplification of lateral deflections due to axial load

This section aims to explain the importance of the Euler’s (P/Pcr ) ratio in design. Rankine
(1866) recognised that the actual failure of structural columns involved an interaction between
elastic and plastic modes of failure. Lateral loads, or geometrical imperfections, lead to the
creation of bending moments in addition to axial loads. Bending moments have to be accom-
panied by stress resultants that diminish the cross-sectional area available for carrying the
axial load, so the failure load PF is always less that the squash load PP .
PF = Failure load i.e. the load at which a column actually fails
PP = Squash load i.e. the yield stress of the material multiplied by the cross-sectional area
Equally, the growth of zones of plastic bending reduces the effective elastic modulus of the
section, thereby reducing the critical load for buckling, so that PF < Pcr . Furthermore these
processes feed on each other, as explained in Horne and Merchant (1965). As the elastic
critical load is approached, all bending effects are magnified.
If lateral loads in the absence of axial load would create a maximum lateral displacement
δo in the critical mode-shape of buckling, then the displacement δ under the same lateral
loads but with a co-existing axial load P is given by Eq. 3.
δ 1
= , (3)
δ0 1 − PPcr
where P is the axial load acting on the pile; Pcr is Euler’s critical load; δ is Deflection of the
pile in the presence of the axial load P δo is deflection of the pile in the absence of the axial
load.
This ratio is often termed as “Buckling Amplification Factor”. The same magnification
factor applies to any initial out-of-line straightness of the pile in the mode shape of potential
buckling. Correspondingly, all curvatures are similarly magnified and so are the bending
strains induced in the column by its lateral loads or eccentricities. The progression towards
plastic bending failure is accelerated as axial loads approach the elastic critical load (Pcr ). Not
only do axial loads induce extra bending moments (P- effects), but the full plastic bending
resistance cannot be mobilized due to the fact that part of the pile section is required to carry
the axial loads. Equation 3 indicates that for a column carrying an axial load of half its Euler
load, that lateral displacements and therefore bending moments would be 1/(1 − 0.5) or
100% bigger than those calculated ignoring axial load effects. This is important if significant
lateral loads must also be carried by the pile. Equation 3 is sketched in Fig. 4g. In the figure
exact equations derived by Halabe and Jain (1993) for different boundary conditions of piles

123
428 Bull Earthquake Eng (2008) 6:407–446

are sketched. The figure shows as the (P/Pcr ) ratio approaches 1, the amplification of lateral
deflections tends to infinity.

5.6 Applicability of “Buckling Amplification Factor” to piles in liquefiable soils

Piles are likely to have greater out-of-straightness, more significant yielding during con-
struction, more deterioration during life, and more uncertainty in their lateral loading during
earthquakes, than is the case with ordinary structural columns in buildings. Furthermore,
the consequences of plastic buckling failure is invariably catastrophic. The creation of a full
plastic mechanism due to the amplification of lateral load effects would lead immediately to
the Euler buckling load reducing to zero, so that the superstructure must collapse suddenly
and violently unless its self-weight can immediately be transferred to foundation elements
that can still carry load. It would therefore be unwise to use a factor of safety less than 3
against the Euler load of a pile.
The question whether earthquake-induced liquefaction can actually eliminate sufficient
lateral soil stiffness, for a long enough period of time, to promote pile buckling in the manner
set out above has been first explored by Bhattacharya (2003), Bhattacharya et al. (2004)
through dynamic centrifuge testing. His experiments showed that a single pile can become
laterally unstable under the axial load alone when the soil surrounding the pile liquefies. More
recently, Knappett and Madabhushi (2005) has shown that a pile group can also collapse by
buckling instability when the soil no longer offers support.

5.7 Interaction formula to account for plastic buckling

An interaction formula to estimate the actual failure load PF in plastic buckling, follow-
ing Rankine’s suggestion, and supported by subsequent research on the collapse of building
frames, was recommended by Horne and Merchant (1965), see Eq. 4.
1 1 1
= + (4)
PF Pcr PP
Structural engineers apply one or other interaction formula of this type to columns, i.e.,
to members in a building frame whose main purpose is the transmission of axial load. Its
application to foundation piles is therefore equally essential so long as it is accepted that soil
liquefaction removes all effective lateral bracing.
Figure 4h shows the plot of 12 concrete piles mentioned in Table 3. The piles are assumed
to be of M25 grade concrete, with a characteristic strength of 25 MPa. In the plot, three
well-defined lines are drawn. These are:
(a) Yield stress line (σy = 11.2 MPa) taken as the design crushing value,
(b) Euler’s curve for σcr , given by Eq. 5 which is the elastic stability limit obtained by
substituting the value of I (Second Moment of Area) from Eq. 1 into Eq. 2 and noting
that σcr is the critical stress given by (Pcr /A).
(c) A curve for σf drawn using Rankine’s formula (1866) shown by Eq. 6, which is essen-
tially Eq. 4. This design curve mediates the transition between strength and stability.
Many other similar curves, such as Perry-Robertson (1886 & 1925), can equally be
used.
Pcr π2
σcr = =  2 E (5)
A Leff
rmin

123
Bull Earthquake Eng (2008) 6:407–446 429

1 1 1
= + , (6)
σf σcr σy
where σy is yield stress of the material and σcr is the elastic critical stress as calculated by
Eq. 5, leading to an estimate of the combined failure stress σf .
The following conclusions can be drawn:
1. The evidence of field studies in Fig. 4e, f, and h is therefore consistent with the hypothesis
that pile failure in liquefied soils is similar to the plastic buckling of structural columns.
2. The lateral support offered to the pile by the soil prior to an earthquake seems capable
of being removed during liquefaction.
3. It may therefore not be relevant to invoke lateral spreading of the soil as the cause of
pile collapse since piles can collapse by plastic buckling as soon as sufficient soil has
liquefied, and before lateral spreading takes place. In addition, the lateral loading applied
by the soil during spreading will only worsen the pile instability.
Nevertheless, doubt may still be expressed on account of the assumption that the pile axial
loads (P) in the case histories could be approximated by the permissible axial loads as plotted
in Fig. 4f. This aspect is further explored by the study of the pile failure of Showa Bridge
during the Niigata earthquake of 1964 in the next section. More details of this study can be
found in Bhattacharya et al. (2005). The actual axial load in the pile has been calculated by
considering the dead load of the superstructure.

5.8 Revisit of the failure of Showa Bridge

5.8.1 Axial load at failure

The dead load on each pile can be deduced from the configuration of the Showa Bridge
deck. Information on the dimensions of the girders, span and the bridge type is obtained
from Iwasaki (1984). Reasonable assumptions are made for the missing data. A schematic
diagram of the deck is shown in Fig. 4i. The bridge has a total length of 303.9 m (13.75 m +
10@ 27.64 m + 13.75 m) and width of 24 m. As mentioned earlier, the superstructure of the
bridge consists of 12 composite steel simple span girders. There are 9 piles in a row sharing
the load of the superstructure. Table 4 shows the estimate of the total dead load for each span.
It is assumed in the analysis that the load of the deck is shared equally by the 9 piles. The
dead load per pile is approximately 740 kN. If the live load due to vehicles were added, the
total axial load P at the time of failure would be roughly 800 kN.
An allowable vertical load calculated on the basis of soil resistance (SPT N values) was
found to be 965 kN, Bhattacharya (2003), which is viewed entirely reasonable in isolation.

Table 4 Dead load calculation for Showa Bridge

Item Details Load (kN)

Slab and asphalt top 24.8 m × 27.64 m × (0.2 + 0.05) m × 25 kN/m3 4146
Crash barrier 2 × 1.5 m × 0.1 m × 27.64 × 25 kN/m3 207.3
Kerb 2 m × 0.15 m × 27.64 m × 25 kN/m3 207
9 steel girders 27.64 m × 9 × 5.2 kN/m 1294
Stiffeners, bolts 30% of girder weight 388
Bottom Girder 0.7 m × 1.0 m × 24 m × 25 kN/m3 420
Total 6662
Load per pile = 6662 kN/9 740

123
430 Bull Earthquake Eng (2008) 6:407–446

5.8.2 Elastic critical load

The length of the pile in the liquefiable soil zone is about 9 m according to Fig. 3f, but the
underlying denser sand can be estimated to require a further 2 m of embedment to provide
moment fixity in the event of liquefaction. The additional length of the structural column
extending the pile and passing through water and air to support the bridge deck is 9 m. In
addition, the bridge dead weight can be regarded as being located 1 m above the top of
the column. Thus the equivalent unsupported length of the cantilever pile/column during
liquefaction is Lo ≈ 21 m. From the buckled shape (shown as the original position in Fig. 3f),
it is clear that the pile had a fixed-free boundary condition in the fashion of Case 6 in Fig.
4d, and hence the effective buckling length Leff = 2Lo ≈ 42 m.
It should perhaps be mentioned that the elastic stiffness of the pile has ignored any con-
tribution from the soil inside, as is conventional in foundation engineering. The pile had two
sections: The lower 8 m of the unsupported length is a 9 mm thick tube and the upper 13 m is
a 16 mm tube with a welded connection about 3 m below the river bed as shown in Fig. 3f.
Timoshenko and Gere (1961) article 2.14 provide an analytical solution for estimating the
buckling load of a column of varying section. The solution proceeds in terms of an upper
length L1 with bending stiffness EI1 and a lower length L2 with bending stiffness EI2 . Cor-
responding buckling parameters k1 and k2 are defined such that k21 = Pcr /EI1 , k22 = Pcr /EI2 ,
where Pcr is the axial load in both sections at the elastic critical load (recalling that there
is no skin friction to be applied in liquefied soil). The solution is found in the form of the
transcendental equation:

tan k1 L1 tan k2 L2 = k1 /k2 (7)

On substituting
L1 = 13 m,
EI1 = 275 MNm2 ,
L2 = 8 m,
EI2 = 160 MNm2 ,

A solution is found by iteration for which k1 = 0.0612 m−1 and k2 = 0.0803 m−1 at
Pcr = 1030 kN. The comparative buckling load if the lower section applied throughout,
given an effective buckling length Leff = 21 × 2 = 42 m, would have been 895 kN as shown
below.

π
Moment of inertia of the pile (I) = (0.6094 − 0.5914 )m4 = 7.63 × 10−4 m4
64

Effective length of pile (Leff ) = 2 × 21 m = 42 m


2 π2
Buckling load of pile (Pcr ) = Lπ2 EI = 422 m2
× 210 GPa × 7.63 × 10−4 m4 = 895 kN
eff
The factor 1.72 enhancement of stiffness in the upper section has resulted only in a factor 1.15
on the elastic buckling load, since it is the lower, lighter section that has the greater curvature
in the cantilever buckling mode. Thus the buckling load of the pile Pcr = 895 × 1.15 =
1030 kN.
Therefore (P/Pcr ) of the pile at failure is 0.77 and therefore the amplification of lateral
deflections due to lateral loads will be approximately 4.5 times, following Fig. 4g and Eq. 3.

123
Bull Earthquake Eng (2008) 6:407–446 431

5.8.3 Failure loads of Showa Bridge piles

Figure 4j shows a typical variation with depth of Euler’s elastic critical load (Pcr ) for
the Showa Bridge pile that receives no lateral bracing from surrounding soil. The pile
has an external diameter of 609 mm and a wall thickness of 9 mm and is passing through
liquefiable soil. The lower pile section of the pile is chosen. The elastic critical load (Pcr ) is a
conventional static estimate, based on the removal of soil lateral stiffness due to soil lique-
faction, the pile being fixed into a rock-socket at the base but free to rotate at the top. Also
shown is the plastic squash load (PP ) that would separately apply if the pile section were
compressed in the absence of bending and buckling; a plastic yield stress of 500 MPa was
adopted.
From Fig. 4j, the Showa Bridge pile would be expected by Rankine’s formula (PF in
the Figure) to fail by plastic buckling under an axial load of 1000 kN, for example, if its
unsupported length exceeded about 18 m. However, structural engineers invariably demand
a high factor of safety against linear elastic buckling to allow for the additional effects of
lateral loading as described in the earlier section. It should therefore be considered essential
to place the design point in Fig. 4j well below the Pcr /2 line, which lies below Rankine’s
failure load PF for long columns under axial load alone. If the tubular pile of Fig. 4j were
used to support an axial load of 1000 kN over an unsupported length of 14 m, and under the
influence of significant lateral loads, it would be likely to collapse.
Referring to Showa Bridge piles, Fig. 4j, a 0.6 m diameter pile fixed at its base and free
at the top has rmin = 0.21 m and Leff = 2 Lo. A slenderness ratio of 50 (the magic slen-
derness ratio in Fig. 4e and h) therefore translates into a free length Lo ≈ 5 m for which
PF < Pcr /2, as expected. Correspondingly, the 0.6 m diameter Showa Bridge pile analysed
in Fig. 4j should not be designed to carry 1000 kN over an unsupported length in excess
of about 11 m. It may however be seen from there that an 800 kN axial load should ideally
buckle elastically at Lo ≈ 22 m.
It may therefore be concluded that the failure of Showa Bridge, which could not be
explained by lateral spreading can be explained by buckling instability.

6 On the strength of liquefied soil: can the resistance of liquefied soil prevent the pile
from buckling?

It is necessary to mention the response of liquefied soil in the context of buckling of a slender
pile. In Seed (1987), H.B. Seed proposed a relationship between liquefied shear strength and
(N1 )60 which was later updated by Seed and Harder (1990). Table 5 lists the value for various
values of N. This relationship was derived by back calculating the shear strengths from a
number of liquefaction flow failures and lateral spreads. This remains the state-of-the-art
practice to estimate the liquefied shear strength for use in post-triggering stability analysis.
More recently, Pillai and Salgado (1994), Olson and Stark (2002) and Olson (2006) pro-
posed some updated relations such as:

Olson and Stark (2002)

Su (Liq)
 = 0.03 + 0.0075[(N1 )60 ] ± 0.03 (N1 )60 in blows/0.3 m
σν0

123
432 Bull Earthquake Eng (2008) 6:407–446

Table 5 Recommended shear strength of liquefied soil, from Seed and Harder (1990)

(N1 )60 Shear Strength (kPa)


Upper bound Lower bound

4 11 0
8 20 2
12 32 10
16 – 25
(N1 )60 : Corrected blow count which would have been obtained under an effective overburden pressure of
98 kPa when the driving energy in the drill rods is 60% of the theoretical free-fall energy of the SPT hammer

Pillai and Salgado (1994)

Su (Liq)
 = 0.21
σν0
It is interesting to note the buckling considerations in pile design as suggested by
Eurocode 7 (1997):

“Slender piles passing through water or thick deposits of very weak soil need to be
checked against buckling. This check is not normally necessary when piles are com-
pletely embedded in the ground unless the characteristic undrained shear strength is
less than 15 kPa”.

Therefore, it seems that the consideration of buckling of pile in liquefiable soil is important
as the shear strength of liquefied soil drops down to very low value at full liquefaction.

6.1 Behaviour of post liquefied sand

Figure 4k shows the stress-strain behaviour of liquefied soil as observed in a multi-stage


triaxial test. The soil was initially liquefied by applying cyclic loading and then mono-
tonic shear was applied. The figure shows that liquefied soil offers no resistance for a
particular amount of strain and then starts offering resistance. This is due to suppressed
dilation.
Based on the above, it is obvious that the shear strength of liquefied ground can be very
small and often below 15–20 kPa. Under these conditions the pile foundation can start to
suffer buckling instability. It is also known that liquefied ground when subjected to mono-
tonic shearing can dilate when its stress state crosses the Phase Transformation Line (Ishihara
1993, 1996). This dilation can translate itself into temporary strengthening of the liquefied
ground until pore fluid from neighbouring free field migrates into the region surrounding the
pile. However in order for this dilation to occur the soil needs to suffer significant amount
of shear strains (see Fig. 4k, for example). As buckling of piles is essentially a structural
instability, the pile failure in this mode can occur well before the soil can regain some of
its strength due to dilation. It was observed first by Bhattacharya (2003) for single piles and
later by Knappett and Madabhushi (2005) for pile groups that the location of plastic hinge
formation is effected by the monotonic shearing and subsequent strengthening of the lique-
fied soil. This, however, does not effect the mode of failure i.e. the pile foundations do suffer
buckling failure in liquefied ground.

123
Bull Earthquake Eng (2008) 6:407–446 433

7 Major research on the mechanisms of pile failure

Following the large-scale damage to piled structures, such as buildings, bridges and port
facilities, during the 1995 Kobe earthquake research into the failure mechanisms of
pile foundations was intensified. Permanent lateral deformation or lateral spreading is
reported to be the main source of distress to piles, for example Dobry and Abdoun (2001),
Tokimatsu (1999); Hamada (1992a), Hamada (1992b). Hamada (2000) concludes that
permanent displacement of non-liquefied soil overlying the liquefied soil is a governing
factor for pile damage. Based on the assumption that lateral spreading is the cause of failure,
research into this pile failure mechanism has been conducted by various researchers, such as
Sato et al. (2001), Takahashi et al. 2002, Haigh (2002), Berrill et al. (2001), Tokimatsu et al.
(2001). A summary of some research publications in the aftermath of 1995 Kobe earthquake
is presented in Table 6. A noteworthy paper by O’Rourke et al. (1994) is also included.

8 The story so far about the mechanisms of failure

The major conclusions based on 11 years of research following the 1995 Kobe earthquake
can be summarized as follows:

1. Earthquakes induce lateral loads on the pile mainly by two ways:


(a) Inertia load which acts at the pile head. This is transferred from the superstruc-
ture to the pile. This force is cyclic in nature and is predominantly a function of
frequency of the super structure and the input motion. This load varies during the
time of earthquake as the stiffness of the soil changes. This is normally maximum
in the initial part of the shaking but before liquefaction. Once the soil liquefies, this
force at the pile head reduces.
(b) Kinematics force acts along the embedded length of the pile till the depth to which
the soil has liquefied. This force is due to the motion of the liquefied soil. If there
is any non-liquefied crust above the liquefied soil, the force applied to the pile
due to this crust will be considerably higher. This force is a function of the slope
of the ground and the type of soil. Physical modelling and case histories provide
example of kinematic forces acting during shaking, after shaking and sometimes
a mixture of both. The timing is related to several phenomena such as duration
of the earthquake shaking, redistribution and dissipation of pore water pressures,
cracking, piping etc. This force will be monotonic if only the kinematic force is
acting i.e. shaking has stopped. But the force will normally be “weakly” cyclic if
both inertia and kinematic force acts together. Most of the current research efforts
are concentrated on estimating or predicting this kinematic load, see for example
Towahata (2006).
2. Under cyclic loading, pore pressures build up in the soil and its stiffness degrades. The
resistance to the movement of the pile will also degrade. The resistance to the pile for a
particular amount of deformation is expressed as p–y soil springs (displacement based
method as is explained later on). These springs degrade with liquefaction. Liu and Dobry
(1995) used the results of a series of centrifuge tests to derive a correlation between the
degradation coefficient and the pore pressure ratio (ru ) as shown in Fig. 5a. ru of 1.0
signifies full liquefaction i.e. 100% pore pressure. These degradation coefficients can be
multiplied to the static p–y relations to obtain the dynamic p–y response.

123
434

Table 6 Research work on mechanism of pile failure during (1995–2006)

SL No Research group and reference Aspect of the problem investigated Conclusions Remarks

123
1 Sesov et al. (2006) Group effects of pile without any axial Specifications for Highway Bridges The drag forces on the piles were
load acting on the pile. Seismic Design (JRA) reasonably measured. Axial load was not
predict the total lateral force which applied on the pile.
is acting on the pile group due to
lateral spreading of liquefied soil.
The total lateral force cannot be
uniformly distributed among the
piles in the pile group.
2 He et al. (2006) Lateral pressure exerted by liquefied The pressure of the liquefied soil on Only kinematic effects of liquefied
soil in a large scale shaking table the pile is in the range of 20 to soil on the pile are quantified. No
tests. 40 kPa. axial load was applied on the piles.
3 Knappett and Madabhushi Buckling Instability of pile groups due Axially loaded end-bearing pile group Buckling can occur in pile groups.
(2005) to axial loads in a centrifuge test. can suffer buckling instability.
Finite element simulation of buck- Partial liquefaction can also lead
ling instability is also carried out to collapse of imperfect pile
using ABAQUS. groups.
4 Tokimatsu and Suzuki (2005) Method to combine Inertia and Kine- This is dependent on the time period of Effects of axial load are not taken into
matic load acting on the pile. the superstructure and that of the consideration.
ground.
5 Rollins et al. (2005) Brigham Lateral load tests on full scale 3 × 3 p–y curves for liquefied sand was The soil surrounding the pile was
Young University pile group to study the pile-soil- developed. Limit equilibrium liquefied by blast.
pile method can be used for analysis.
interaction effects
6 Haigh and Madabhushi (2005) Transient lateral forces experienced Conclusions arrived through centri-
University of Cambridge by the pile during lateral spread JRA (1996) code (Figure 1.6) is fuge tests using the stress cells
down a slope. Rigid and flexible un-conservative by 300% in tran- measurements. Near field pore
piles were used. No axial load was sient phase but gives reasonable pressures were also measured.
applied. predictions at residual values.
7 Berrill et al. (2001) University Good performance of pile foundations The chief threat to piled foundations Limit equilibrium analysis was used.
of Canterbury, New of Landing bridge. comes from the non-liquefied
Zealand crust and not from the drag force
of the liquefied soil.
Bull Earthquake Eng (2008) 6:407–446
Table 6 continued

SL No Research group and reference Aspect of the problem investigated Conclusions Remarks

8 Sato et al. (2001), National Lateral forces experienced by a pile The forces predicted by JRA (1996) Conclusions arrived through
Research Institute for foundation near a quay wall are over-conservative. centrifuge tests using stress
Earth science and disaster during lateral spread. cells measurements. The
prevention, Japan measurements showed no trends.
9 Tokimatsu et al. (2001) Tokyo The p–y behaviour (i.e. relation If the pile pushes the soil the subgrade Shaking table tests were carried out
Institute of Technology, between subgrade reaction and reaction is co-related with relative using a large-scale laminar box.
Japan relative displacement between soil soil pile displacement, and if soil
and pile) during soil liquefaction. pushes the pile the subgrade reac-
Bull Earthquake Eng (2008) 6:407–446

tion can be co-related with relative


velocity between pile and soil.
10 Ramos et al. (2000) R.P.I (New Effect of superstructure’s horizontal Simple limit equilibrium approach Centrifuge tests were carried out on
York) stiffness on the bending moments using the pressure distribution end-bearing piles in absence of
induced on the pile by lateral suggested by Dobry and Abdoun axial load. The bending moments
spreading. (1998) gives a good prediction of in the pile were measured using a
the bending moment in the pile. pair of strain gauges.
11 Hamada (2000) Waseda Characterisation of the external forces The force from flowing liquefied soil He notes that the deformation of the
University, Japan from the flowing liquefied soil and on a model pile can be estimated model pile can be simulated by a
non-liquefied crust on model piles as a drag force against a cylindri- beam and soil spring model where
under 1-g condition. 26 mm poly- cal object in a viscous flow. The the displacement of the soil is
carbonate pipe was used as model external force from non-liquefied forced into the pile through the soil
pile. soil overlying the flowing lique- spring.
fied soil governs the deformation
of the pile.
12 Wilson et al. (2000) University Lateral p–y resistance of liquefied The lateral p–y resistance of liquefied They note that there is considerable
of California (Davis) soils. soil is a complex phenomenon and uncertainty in any simplified rep-
Centrifuge tests were carried out on is significantly affected by relative resentation of p–y characteristics
single pile and pile-group sup- density, cyclic degradation, excess of liquefied soil and this uncer-
ported structures in liquefied sand. pore pressures, phase transforma- tainty must be allowed for in
tion behaviour, prior displacement design.
history and loading rate.

123
435
436

Table 6 continued

123
SL No Research group and reference Aspect of the problem investigated. Conclusions Remarks

13 Goh and O’Rourke (1999) Development of numerical code to The strain softening p–y model pro- Dynamic effects are not included.
Cornell University calibrate the centrifuge test results vides excellent predictions of Haigh (2002) extended the above
of Abdoun (1997). The model uses the measured peak and residual model to a pseudo dynamic p–y–u
a tri-linear strain softening p–y moments. analysis where u is the pore pres-
curves. The curves are obtained The computed soil pressure distri- sure rise.
from FLAC analysis of a pile bution agrees well with the JRA
being displaced through a cohe- (1996) code.
sive material. The undrained shear
strength in the curve initially rises
to a peak value and then falls line-
arly with plastic deviatoric strain
until a residual shear strength is
achieved.
14 Dobry and Abdoun (1998) Pressure distribution acting on a pile The maximum bending moment in a Limit equilibrium analysis was used
R.P.I (New York) during lateral spreading. The pres- pile could be obtained by applying to fit the bending moment data
sure distribution was back calcu- an inverted triangular distribution of instrumented piles obtained in
lated from the bending moments. having a value of 17.7 kPa at the centrifuge tests.
surface to zero at 6 m depth.
15 Abdoun (1997) RPI (New Bending moments acting on single Non-liquefied crust exerts approxi- A series of centrifuge tests were car-
York) piles and pile groups in layered mately passive pressure to a pile ried out in two-layer system and
soils. foundation. three-layer system.
16 Liu and Dobry (1995) Stiffness degradation of liquefied soil They derived a dimensionless degra- A series of centrifuge tests were car-
due to excess pore pressure rise so dation coefficient against ru (ratio ried out.
as to develop a dynamic p–y curve of excess pore pressure to that
for design of pile foundations. causing full liquefaction). This
coefficient can be multiplied to
p–y curves for static tests to gen-
erate dynamic p–y curves.
Bull Earthquake Eng (2008) 6:407–446
Table 6 continued

SL No Research group and reference Aspect of the problem investigated. Conclusions Remarks

17 O’Rourke et al. (1994) Analytical studies of pile response to Three failure mechanisms viz. The pile was modelled as a series
Cornell University lateral spread using the computer excessive bending, buckling and of beam element whereas the sur-
code B-STRUCT. Dimensionless soil flow were recognised under rounding soil is modelled as trans-
plots were developed. lateral spread condition. Each of verse and longitudinal bilinear
the mechanisms is a function of spring-slider elements.
relative stiffness between soil and
pile and the axial load carried by
the pile.
Bull Earthquake Eng (2008) 6:407–446

123
437
438 Bull Earthquake Eng (2008) 6:407–446

Fig. 5 (a) Pore pressure ratio ru versus Degradation Parameter Cu . (b) Combination of inertial force and
ground displacement (Tokimatsu et al. (2005)). For Case (a) the moment M is simply the addition of the two
components. (c) Stiffness degradation for type-2 soil profiles (Ishihara and Cubrinovski 1998). (d) Stiffness
degradation for type-1 soil profiles (Ishihara 1997). (e) Effect of liquefaction on p–y behaviour, Wilson et al.
(2000). In the above figure (a) denotes p–y loops at a depth of 2D, (b) denotes p–y loops at a depth of 3D and
(c) denotes p–y loops at a depth of 4D. Dashed line represents API (2003) p–y curves. (f) Strain softening
p–y curves (Goh and O’Rourke 1999). In the axes, D is pile diameter, y is the pile displacement and sup is the
undrained strength of soil

123
Bull Earthquake Eng (2008) 6:407–446 439

Fig. 5 continued

123
440 Bull Earthquake Eng (2008) 6:407–446

3. Most often, piled foundations are used to support a structure in soft grounds as the soil
cannot bear the load. Piles transfer the axial loads of the superstructure to the deep
stronger strata by the shear generated along the shaft and the end bearing. Buckling of
piles under such axial load is not normally an issue as the confining pressure of the
soil is sufficient to provide lateral bracings. During earthquake liquefaction, the soil no
longer provides sufficient lateral support and a pile may become unstable. The mech-
anism is buckling instability. Buckling is sensitive to imperfections such as out-of-line
straightness, lateral loads, eccentricity of axial loads. Basically, a pile should be treated
as unsupported in the liquefiable zone.
4. There is a requirement of combining the effects of Inertial and Kinematic loads during
the design as mentioned in the first point above. Tokimatsu et al. (2005) undertook large
scale shaking table tests of pile groups in dry and saturated sands and suggested that
the phasing depends on the relationship between the ground (Tg ) and the structure (Tb )
natural periods as follows:.

(a) Tb > Tg — Kinematic forces tend to be out of phase with the inertial loads and
peak pile stresses will occur at an interim point.
(b) Tb < Tg — Kinematic and inertial forces tend to be in phase, and peak pile stresses
occur when both effects are a maximum.
5. In order to estimate the peak bending moment due to the combined inertial and kine-
matic loads, Tokimatsu et al. (2005), Tokimatsu and Suzuki (2005) suggests the following
method the two effects as follows:

(a) Tb > Tg —The peak pile bending moment in the pile can be estimated by the SRSS
[Square Root of the Sum of Squares] of the individual moments due to Inertia and
Kinematic loads.
(b) Tb < Tg —The peak bending moment in the pile can be computed by the algebraic
addition of the inertia and kinematic component.

9 Design approaches

The current design methods are aimed to protect the pile against bending failure due to lateral
loads such as inertia and lateral spreading. There are two methods used in practice:
1. Force based method or “Limit Equilibrium method”: This is a direct straight forward
method where the lateral pressure acting on the pile is estimated and then the response
of the pile is evaluated i.e. checked against pile yielding and the allowable deflection.
2. Displacement based method: In this method, the free-field ground displacement is evalu-
ated first and then this displacement profile is applied to the pile. In Japan, this is termed
as “Seismic Deformation method”.

10 Force based methods or limit equilibrium method

In this method, the response of the pile is evaluated based on the lateral soil pressure acting
on the pile. The following pressure profiles are available:
1. JRA (1996) Approach, explained in the earlier sections, see Fig. 3b.

123
Bull Earthquake Eng (2008) 6:407–446 441

2. Rensselaer Polytechnic Institute i.e. Abdoun (1997) approach and their revised He et al.
(2006) approach: In the 1997 approach it was proposed that the liquefied soil imposes
a uniform pressure of 10 kPa on the piles. This was to fit the centrifuge test data. In the
2006 approach, it is proposed that liquefied soil offers a uniform pressure of 20 kPa to
40 kPa.

10.1 Limit equilibrium method

Dobry et al. (2003) presented a simplified limit equilibrium method for computing maximum
bending moments in a pile, given by Eq. 8.
Mmax = (0.5Ap · Hp + Ac · Hc )pL , (8)
where Ap and Hp are the area and length of the pile within the liquefied layer, Ac and Hc are
the area and depth of the pile cap and PL is the limiting liquefied soil pressure.

10.2 Displacement based method or p–y method or seismic deformation method

For the analysis of laterally loaded piles various analytical models have been proposed. The
most commonly used model for predicting the non-linear behaviour of soil is using p–y
curves (API 2003). The p–y curves are required for solving the basic differential equation
for a laterally loaded pile, shown by Eq. 9
d 4y
EI + Es · y = F, (9)
dx 4
where y is the lateral deflection; x is distance along the pile; EI is flexural stiffness of pile;
ES is soil modulus; F is applied force per unit length of the pile.

The equation is most easily solved numerically, iterating to satisfy the boundary condition
at the top of the pile for equilibrium and compatibility. The concept of developing p–y curves
giving soil resistance p = ES · y as a function of pile deflection y, is the most preferred
method of solving Eq. 9. The solution gives pile deflection, pile rotation, bending moment,
shear and soil reaction for any load capable of being sustained by the pile.
The generation of p–y curves is based on field tests carried out in Texas; see for example
Matlock (1970), Reese et al. (1974, 1975). From these field tests, theoretically based, empir-
ical formulae to give good correlations with field data have been established for static and
cyclic loading. The p–y curves are generated based on the following parameters:
1. p = Lateral soil resistance
2. pu = Ultimate lateral soil resistance
3. y = Lateral pile deflection
4. yc = 2.5ε50 d
5. ε50 = Characteristic strain at 50% of failure stress in undrained tests
6. d = Pile diameter

10.3 p–y approach in AIJ (Architectural Institute of Japan) code

In earthquake engineering, Eq. 9 has been adopted in the following way, see AIJ (1988,
2001):
d 4y
EI = −kh · Bp (y − yg ), (10)
dz4

123
442 Bull Earthquake Eng (2008) 6:407–446

where Z is the depth; y is pile displacement; yg is ground displacement; EI is flexural rigidity


of the pile.
The coefficient of subgrade reaction kh is given by Eq. 11 as follows:


kh = kh1  , (11)
 
1 +  yy1r 

where Kh1 = 80 · E0 · B0−0.75 , E0 is the 0.7 N in the units of MN/m2 and N is the SPT value;
β = Scaling factor for liquefied soil. The value is generally 1 for non-liquefied sand and 0.1
for liquefied sand; B0 is pile diameter in cm; yr is relative displacement between pile and the
soil = (y − yg ); y1 is reference value of yr . This is often used as 1% of pile diameter.

10.4 Typical values of β (Scaling factor for liquefied soil)

Ishihara (1997) also suggested a multiplicative factor denoted by β to account for stiffness
degradation due to liquefaction, such that the p–y curves are given by Eq. 12.

p = βk(y, z)y (12)

For level ground problems, β is typically taken to be 0.1 (10%) at ru = 1.0.


For laterally spreading soils with non-liquefiable crusts, the variation of β with norma-
lised relative displacement of the layer has been estimated from back analyses by Ishihara
and Cubrinovski (1998), see Fig. 5c.
For laterally spreading soil without non-liquefiable crusts, the variation of β with norma-
lised relative displacement of the layer has been estimated from back analyses by Ishihara
(1997), see Fig. 5d.

10.5 P–y method based on U.C. Davis

Wilson et al. (2000) studied the p–y behaviour of liquefying sand in soil-pile-structure exper-
iments with single piles. Centrifuge tests were undertaken with liquefiable Nevada sand at
relative densities of RD = 35% and RD = 55%. The lateral soil resistance for the loose sand
(RD = 35%) was small, even for large relative lateral displacement (y/D ≈ 0.1). In contrast,
the medium dense sand exhibited a peak resistance followed by a softening as the excess
pore pressure ratio increased to ru ≈ 1.0. These features are shown in Fig. 5e.

10.6 Goh and O’Rourke (1999)

Goh and O’Rourke (1999) presented a strain softening p–y curve shown in Fig. 5f, corre-
sponding to a soil whose undrained shear strength reduces with strain to a residual value at
larger displacements. This type of p–y curve can be used to model the response of medium-
dense to dense soils, but not loose soils at low stress levels, which do not generally show such
a peaked response. This curve was used to match the centrifuge results of Abdoun (1997) for
a laterally spreading soil profile.
The above concept of p–y curves was modified to include the effect of excess pore pressure
generation by considering the p–y–u space by Haigh and Madabhushi (2005). In this paper
they propose normalising the p–y curves with the excess pore pressure magnitude to account
for the liquefaction effects on the pile behaviour.

123
Bull Earthquake Eng (2008) 6:407–446 443

11 Conclusions

1. Two different theories of pile failure in liquefiable soils have been described citing exam-
ples and explanation behind the theories. The various methods for seismic pile design
in liquefiable areas have also been reviewed. The current methods for pile design are
based on a bending mechanism and the piles are treated as laterally loaded beams. The
most recent Japanese Highway Code (JRA 2002) which is one of the most advanced
codes has been reviewed. The recommendations in this code along with the background
theory behind such recommendations have been critically analysed. The limitations of
this code have been highlighted by taking the example of well-documented case history
of the failure of Showa Bridge during the 1964 Niigata earthquake. It has been shown
that even though the design of piles would satisfy the current JRA code of practice, the
piles actually failed during the 1964 Niigata earthquake.
2. The large scale damage to piled foundations during the 1995 Kobe earthquake has
intensified the research on the mechanism of failure of piled foundations in liquefiable
soils around the globe. The summary of the research carried out at Cornell University
(USA), Rensselaer Polytechnic Institute (USA), Tokyo Institute of Technology (Japan),
University of California (Davies), University of Cambridge (UK), University of Can-
terbury (NZ), University of Oxford (UK), University of Tokyo (Japan), NIED (Japan)
and Waseda University (Japan) have been reviewed. Most of the research carried out
concentrates on the bending failure of the pile. The theory based on bending failure
assumes that failure of piles in liquefied soils is by lateral spreading of the soil pushing
the pile over. Therefore, a mildly sloping ground is necessary for a building to collapse.
In other words, a structure can only collapse in laterally spreading ground.
3. However, it has been shown that the failure patterns of buildings in level ground i.e. in
the absence of lateral spreading is similar to the failures observed in laterally spreading
ground. Moreover, hinges formation was observed in excavated piles. All current design
codes produce a large margin of safety against hinge formation (using partial factors on
load and material stress), yet occurrences of pile failure due to liquefaction are abun-
dant. Bhattacharya (2003) has shown that in some cases, the overall factor against plastic
hinging at collapse has been of the order of 4 to 8. This is strong evidence that there
is another mechanism, which the code does not consider, governing these failures. This
paper suggests that this missing element is the removal of the lateral support the liquefied
soil gives to the pile, making it susceptible to buckling under the axial load. The theory
based on buckling failure therefore assumes that when the soil liquefies the reduction in
the supporting lateral stiffness of the soil can cause a buckling failure of the pile which
can be initiated by lateral spreading or inertia or any in-built imperfections.
4. Buckling of piles is not new to civil engineers. Codes such as API (2003) or Eurocode
does consider column buckling, but only for soils having low shear strength i.e. soft clay
where the undrained shear strength is less than 15 kPa. In contrast, this paper points out
that fully embedded end-bearing piles passing through loose to medium dense sand can
buckle under the axial load, if the soil surrounding the pile liquefies.
5. Buckling and bending requires different approaches in design. Currently, piles in lique-
fiable soils are designed against bending mechanism. Designing against bending would
not necessarily suffice the buckling requirement. To avoid buckling, there is a need for
a minimum diameter of pile depending on the thickness of the liquefiable soils. The
practical implications of this may be far reaching, the most important being a need for
re-assessment of the existing structures in liquefiable soils, and another the need for a
rewriting of the codes to take buckling effects into account.

123
444 Bull Earthquake Eng (2008) 6:407–446

References

Abdoun TH (1997) Modelling of seismically induced lateral spreading of multi-layered soil and its effect on
pile foundations. PhD thesis, Rensselaer Polytechnic Institute, NY
Abdoun TH, Dobry R (2002) Evaluation of pile foundation response to lateral spreading. Soil Dyn Earthq
Eng 22:1051–1058
Abdoun T, Dobry R, O’Rourke TD, Goh SH (2003) Pile response to lateral spreads: centrifuge modeling.
J Geotech Geoenviron Eng 129(10):869–878
AIJ (1988) Architectural Institute of Japan, Recommendations for design of building foundations (in Japanese)
AIJ (2001) Architectural Institute of Japan, Recommendations for design of building foundations (in Japanese)
API (2003) American Petroleum Institute, Recommended practice for planning designing and constructing
fixed offshore platforms
Berrill JB, Christensen SA, Keenan RP, Okada W, Pettinga JR (2001) Case studies of lateral spreading forces
on a piled foundation. Geotechnique 51(6):501–517
Bhattacharya S (2003) Pile instability during earthquake liquefaction. PhD thesis, University of
Cambridge, UK
Bhattacharya S, Madabhushi SPG, Bolton MD (2004) An alternative mechanism of pile failure in liquefiable
deposits during earthquakes. Geotechnique 54(3):203–213
Bhattacharya S, Bolton MD, Madabhushi SPG (2005) A reconsideration of the safety of the piled bridge
foundations in liquefiable soils. Soils Found 45(4):13–26
BTL Committee (2000) Study on liquefaction and lateral spreading in the 1995 Hyogoken-Nambu earth-
quake, Building Research Report No 138. Building Research Institute, Ministry of Construction, Japan
(in Japanese)
Dobry R, Abdoun T (1998) of Conference. Post-triggering response of liquefied sand in the free field and
near foundations. In: Dakoulas P, Yegian M, Holtz RD (eds) Proceedings of geotechnical earthquake
engineering and soil dynamics III. Geotechnical special publication no. 75, ASCE. Seattle, Washington,
August 3–6, 2, pp 270–300
Dobry R, Abdoun T (2001) Recent studies on seismic centrifuge modelling of liquefaction and its effect on
deep foundation. In: Proceedings of the 4th international conference on recent advances in geotechnical
earthquake engineering and soil dynamics and symposium in honour of Professor W. D. Liam Finn. San
Diego, California, March 26–31, 2001
Dobry R, Abdoun T, O’Rourke TD, Goh SH (2003) Single piles in lateral spreads: field bending moment
evaluation. J Geotech Geoenviron Engg ASCE 129(10):879–889
Editorial Committee for the Report on the Hanshin-Awaji Earthquake Disaster (1998) Report on the
Hanshin-Awaji earthquake disaster, Building Series, vol 4, Maruzen (in Japanese)
Eurocode 7 (1997) Geotechnical design, European Committee for Standardization, Brussels
Finn WDL, Fujita N (2002) Piles in liquefiable soils: seismic analysis and design issues. Soil Dyn Earthq Eng
22:731–742
Finn WDL, Thavaraj T (2001) Deep foundations in liquefiable soils: case histories, centrifuge tests and meth-
ods of analysis. In: Proceedings of the 4th international conference on recent advances in geotechnical
earthquake engineering and soil dynamics and symposium in honour of Professor W. D. Liam Finn. San
Diego, California, March 26–31, 2001
Fujii S, Isemoto N, Satou Y, Kaneko O (1998) Investigation and analysis of a pile foundation damaged by
liquefaction during the 1995 Hyogoken-Nambu earthquake. Soils Found., Special issue on geotechnical
aspects of the 17 January 1995 Hyogoken-Nambu Earthquake, No. 2, pp 179–192
Fukuoka M (1966) Damage to civil engineering structures. Soils Found 6(2), pp 45–52
Goh S, O’Rourke TD (1999) Limit state model for soil–pile interaction during lateral spread. In: Proceed-
ings of 7th US–Japan workshop on earthquake resistant design of lifeline facilities and countermeasures
against soil liquefaction. Seattle, WA
Haigh SK (2002) Effects of Earthquake-induced liquefaction on pile foundations in sloping ground. PhD
thesis, University of Cambridge, UK
Haigh SK, Madabhushi SPG (2005) The effects of pile flexibility on pile loading in laterally spreading slopes,
invited paper, ASCE-GI Special publication on simulation and seismic performance of pile foundations
in liquefied and laterally spreading ground. In: Boulanger RW, Tokimatsu K (eds) ASCE Geotechnical
special publication no. 145, ISBN 0-7844-0822-X, pp 24–37
Halabe UB, Jain SK (1993) Amplification factors for piles. J Eng Struct, Butterworth-Heinemann Ltd.
15(2):97–101
Hamada M (1992a) Large ground deformations and their effects on lifelines: 1964 Niigata earthquake.
Case studies of liquefaction and lifelines performance during past earthquake. Technical Report NCEER-

123
Bull Earthquake Eng (2008) 6:407–446 445

92–0001, vol 1. Japanese case studies, National Centre for Earthquake Engineering Research, Buffalo,
NY
Hamada M (1992b) Large ground deformations and their effects on lifelines: 1983 Nihonkai-Chubu earth-
quake. Case studies of liquefaction and lifelines performance during past earthquake. Technical report
NCEER-92–0001, vol 1. Japanese case studies, National Centre for Earthquake Engineering Research,
Buffalo, NY
Hamada M (2000) Performances of foundations against liquefaction-induced permanent ground displace-
ments. In: Proceedings of the 12th world conference on earthquake engineering. Auckland, New Zealand,
paper no 1754
Hamada M, O’Rourke TD (eds) (1992) Case studies of liquefaction and lifeline performance during past
earthquakes, vol 1. Japanese case studies, Technical report NCEER-92–0001
He L, Elgamal A, Abdoun T, Abe A, Dobry R, Meneses J, Sato M, Tokimatsu K (2006) Lateral load on piles
due to liquefaction-induced lateral spreading during one-G shake table experiments. In: Proceedings of
the 100th anniversary earthquake conference commemorating the 1906 San Francisco earthquake, April
18–22, California
Horne MR, Merchant W (1965) The stability of frames. Pergamon
Ishihara K (1993) Rankine lecture: liquefaction and flow failure during earthquakes. Geotechnique 43(3):
351–415
Ishihara K (1996) Soil behaviour in earthquake geotechnics. Oxford Science Publication, Clarendon Press,
Oxford, UK
Ishihara K (1997) Terzaghi oration: geotechnical aspects of the 1995 Kobe earthquake. In: Proceedings of
ICSMFE. Hamburg, pp 2047–2073
Ishihara, K, Cubrinovski M (1998) Problems associated with liquefaction and lateral spreading during earth-
quakes. In: Proceedings of a speciality conference, geotechnical earthquake engineering and soil dynam-
ics III, ASCE geotechnical special publication, no 75, pp 301–312
Iwasaki T (1984) A case history of bridge performance during earthquakes in Japan. In: Prakash S (ed) Key-
note lecture at the international conference on case histories in geotechnical engineering, May 6–11,
University of Missouri—Rolla
JRA (2002) Japanese Road Association, Specification for highway bridges, part V, seismic design
JRA (1996) Japanese Road Association, specification for highway bridges, part V, seismic design
Kawamura S, Nishizawa T, Wada H (1985) Damage to piles due to liquefaction found by excavation twenty
years after earthquake. Nikkei Architecture, 27 May, pp 130–134
Knappett JA, Madabhushi SPG (2005) Modelling of liquefaction-induced instability in pile groups. In:
Boulanger RW, Tokimatsu K (eds) ASCE geotechnical special publication no 145 on seismic perfor-
mance and simulation of pile foundations in liquefied and laterally spreading ground, pp 255–267
Liu L, Dobry R (1995) Effect of liquefaction on lateral response of piles by centrifuge model tests. NCEER
report to FHWA. NCEER Bull 9(1):7–11
Madabhushi SPG, Patel D, Haigh SK, (2005) Geotechnical aspects of the Bhuj Earthquake, Chapter 3, EEFIT
Report, Institution of Structural Engineers, London, UK, ISBN 0901297-372
Matlock H (1970) Correlations for design of laterally loaded piles in soft clay. In: Proceedings of 2nd offshore
technology conference, Houston, paper no OTC 1204
National Research Council (1985) Liquefaction of soils during earthquakes. Report No. CETS-EE-001, com-
mittee on earthquake engineering. National Academy Press, Washington, DC
Olson SM (2006) Liquefaction analysis of Duncan Dam using strength ratios. Can Geotech J 43:484–499
Olson SM, Stark TD (2002) Liquefied strength ratio from liquefaction flow failure case histories. Can Geotech
J 39:629–747
Pillai VS, Salgado FM (1994) Post-liquefaction stability and deformation analysis of Duncan Dam. Can
Geotech J 31:967–978
Ramos R, Abdoun TH, Dobry R (2000) Effects of lateral stiffness of superstructure on bending moments of pile
foundation due to liquefaction induced lateral spreading. In: Proceedings of the 12th world conference
on earthquake engineering, Auckland, New Zealand
Rankine WJM (1866) Useful rules and tables. London
Reese LC, Cox WR, Koop FD (1974) Analysis of laterally loaded piles in sand. Offshore Technology Con-
ference, OTC paper no 2080
Reese LC, Cox WR, Koop FD (1975) Field testing and analysis of laterally loaded piles in stiff clay. Offshore
Technology Conference, OTC paper no 2312
Rollins KM, Gerber TM, Lane JD, Ashford SA (2005) Lateral resistance of a full-scale pile group in liquefied
sand. J Geotech Geoenviron Eng ASCE 131(1):115–125
Rourke TDO, Meyersohn WD, Shiba Y, Chaudhuri D (1994) Evaluation of pile response to liquefaction-
induced lateral spread. In: Proceedings from the 5th US-Japan workshop on earthquake resistant design

123
446 Bull Earthquake Eng (2008) 6:407–446

of lifeline facilities and countermeasures against soil liquefaction: technical report NCEER 94-0026,
National Centre for Earthquake Engineering Research, Buffalo, NY
Sato M, Ogasawara M, Tazoh T (2001) Reproduction of lateral ground displacements and lateral-flow earth
pressures acting on a pile foundations using centrifuge modelling. In: Proceedings of the 4th international
conference on recent advances in geotechnical earthquake engineering and soil dynamics and symposium
in honour of Professor W. D. Liam Finn. San Diego, California, March 26–31, 2001
Seed HB (1987) Design problems in soil liquefaction. J Geotech Eng ASCE 113(8):827–845
Seed RB, Harder LF Jr (1990) SPT-based analysis of cyclic pore pressure generation and undrained resid-
ual strength. In: Proceedings of the H. B. Seed memorial symposium, vol 2. Bi-Tech Publishing Ltd,
pp 351–391
Sesov V, Towahata I, Motamed R (2006) Lateral pressure distribution in a large group pile caused by flow
of liquefied sandy soil. In: Proceedings of the 1st European conference on earthquake engineering and
seismology, Geneva, Switzerland, 3–8th September 2006
Soga K (1997) Chapter 8, Geotechnical aspects of Kobe earthquake. Of EEFIT report, Institution of Structural
Engineers, UK
Takata T, Tada Y, Toshida I, Kuribayashi E (1965) Damage to bridges in Niigata earthquake. Report no 125-5,
Public Works Research Institute (in Japanese)
Tazoh T (2007) Earthquake engineering research on pile foundations with emphasis on piles subjected to large
ground deformations. In: Bhattacharya s (ed) NICEE Special Publications Foundation design in seismic
areas: principles and few applications
Timoshenko SP, Gere JM (1961) Theory of elastic stability. McGraw-Hill book company, New York
Tokimatsu K (1999) Performance of pile foundations in laterally spreading soils. In: Proceedings of the 2nd
international conference on earthquake geotechnical engineering, vol 3. Lisbon, Portugal, June 21–25,
pp 957–964
Tokimatsu K, Asaka Y (1998) Effects of liquefaction-induced ground displacements on pile performance in
the 1995 Hyogeken-Nambu earthquake. Special issue of soils and foundations, pp 163–177, Sep 1998
Tokimatsu K, Suzuki H (2005) Effect of inertial and kinematic interactions on seismic behaviour of pile
foundations based on large shaking table tests. In: Proceedings of the 2nd CUEE conference on urban
earthquake engineering. Tokyo Institute of Technology (Japan)
Tokimatsu K, Mizuno H, Kakurai M (1996) Building damage associated with geotechnical problems. Special
issue of soils and foundations. Japanese Geotechnical Society, Jan 1996, pp 219–234
Tokimatsu K, Suzuki H, Suzuki Y (2001) Back-calculated p–y relation of liquefied soils from large
shaking table tests. In: Proceedings of the 4th international conference on recent advances in geotechnical
earthquake engineering and soil dynamics and symposium in honour of Professor W. D. Liam Finn. San
Diego, California, March 26–31, 2001
Tokimatsu K, Suzuki H, Sato M (2005) Effects of dynamic soil-pile structure interaction on pile stresses.
J Struct Const Eng, Arch Inst Jpn 587:125–132
Tokimatsu K, Oh-oka H, Satake K, Shamoto Y, Asaka Y (1997) Failure and deformation modes of piles due
to liquefaction-induced lateral spreading in the 1995 Hyogoken-Nambu earthquake. J Struct Eng AIJ,
Jpn 495:95–100
Tokimatsu K, Oh-oka H, Satake K, Shamoto Y, Asaka Y (1998) Effects of lateral ground movements on failure
patterns of piles in the 1995 Hyogoken-Nambu earthquake. In: Proceedings of a speciality conference,
geotechnical earthquake engineering and soil dynamics III, ASCE Geotechnical special publication no
75, pp 1175–1186
Wilson DW, Boulanger RW, Kutter BL (2000) Observed seismic lateral resistance of liquefying sand.
J Geotech Eng ASCE 126(10):898–906
Yasuda S, Yoshida N, Kiku H, Adachi K, Gose S (1999) A simplified method to evaluate liquefaction-induced
deformation. In: Seco e Pinto PS (ed) Earthquake geotechnical engineering. © Balkema, Rotterdam,
ISBN 90-5809-1163
Yokoyama K, Tamura K, Matsuo O (1997) Design methods of bridge foundations against soil liquefaction
and liquefaction-induced ground flow. In: 2nd Italy–Japan workshop on seismic design and retrofit of
bridges. Rome, Italy, Feb 27 and 28, 1997
Yoshida N, Hamada M (1990) Damage to foundation piles and deformation pattern of ground due to
liquefaction-induced permanent ground deformation. In: Proceedings of 3rd Japan–US workshop on
earthquake resistant design of lifeline facilities and countermeasures for soil liquefaction, pp 147–161
Yoshida N, Towahata I, Yasuda S, Kanatani M (2005) Discussion to the paper by Bhattacharya et al. (2004)
on an alternative mechanism of pile failure in liquefiable deposits during earthquakes. Geotechnique
55(3):259–263
Yoshida N, Tazoh T, Wakamatsu K, Yasuda S, Towhata I, Nakazawa H, Kiku H (2007) Causes of Showa Bridge
collapse in the 1964 Niigata earthquake based on eye-witness testimony. Soils Found 47(6):1075–1087

123

You might also like