You are on page 1of 11

Technical Note

Liquefaction of soil in the Tokyo Bay area from the 2011


Tohoku (Japan) earthquake
S. Bhattacharya
a,n
, M. Hyodo
b
, K. Goda
a
, T. Tazoh
c
, C.A. Taylor
a
a
University of Bristol, United Kingdom
b
Yamaguchi University, Japan
c
Shimizu Corporation, Japan
a r t i c l e i n f o
Article history:
Received 14 May 2011
Received in revised form
15 June 2011
Accepted 16 June 2011
Available online 7 July 2011
a b s t r a c t
Immediately following the 11th March 2011 M
w
9.0 Tohoku (Japan) earthquake, a eld investigation
was carried out around the Tokyo Bay area. This paper provides rst-hand observations (before or just
at the onset of repair) of widespread liquefaction and the associated effects. Observations related to
uplift of manholes, settlement of ground, performance of buildings and bridges and the effects of
ground improvements are also presented. Recorded ground motions near the Tokyo Bay area were
analysed to understand their key characteristics (large amplitude and long duration). Lessons learnt are
also presented.
Crown Copyright & 2011 Published by Elsevier Ltd. All rights reserved.
1. Introduction
A devastating earthquake of moment magnitude M
w
9.0 struck
the Tohoku and Kanto regions of Japan on 11th March at 2:46 pm.
The earthquake caused great economic loss, loss of life and
tremendous damage to structures and infrastructures. This was
the largest earthquake ever recorded in Japan and one of the ve
most powerful earthquakes in the world since modern record-
keeping began in 1900. According to the Japan Meteorological
Association (JMA), the epicentre was located about 150 km off-
shore from Sendai (largest city in the Tohoku region) and at a
depth of about 24 km (Fig. 1). Intense tremors were felt across
coastal areas of the Tohoku and Kanto regions (67 magnitude
based on JMA seismic intensity scale). Extensive damage was
caused by massive tsunami in many cities and towns along the
coast. At many locations (e.g. Natori, Oofunato and Onagawa),
tsunami heights exceeded 10 m, and sea walls and other coastal
defence systems failed to prevent the disaster. The earthquake
and its associated effects (i.e. tsunami) also initiated the crisis of
the Fukushima Daiichi nuclear power plants. In Tokyo, which is
about 350400 km southwest from the epicentre, a relatively
intense tremor of JMA scale 5 was registered. This earthquake
induced many large aftershocks with M
w
greater than 6.0, which
caused additional damage to structures that survived during the
mainshock. It is expected that the increased seismicity will be
relatively high in the next several months to a few years and
which will impose additional threat to residents in the affected
regions.
This study reports a geotechnical eld investigation in the
Tokyo Bay area that was conducted immediately after the 2011
Tohoku earthquake (1215 March). The studied area (i.e. Shin-
kiba, Urayasu and Tokyo Disneyland; see Fig. 2) was damaged
severely by widespread liquefaction. Organised large-scale eld
investigations by various international agencies were delayed so
that the humanitarian efforts were not disturbed and due to the
radiation problems at the Fukushima Daiichi nuclear power plants.
This survey provides the rst raw details (unrepaired structures
and ground condition) of the effects of the earthquake. Most
importantly, these investigations provide valuable insights on
civil engineering lessons from this catastrophic event and identify
challenges/issues that need to be addressed to mitigate seismic
damage due to future destructive earthquakes.
2. Fault rupture and ground motion characteristics
The M
w
9.0 earthquake occurred off-shore of the Miyagi Pre-
fecture, Japan. This was an interface subduction earthquake with a
low-angle reverse mechanism, and was caused by tectonic move-
ments of the North American (Okhotsk) plate and the Pacic plate.
The relative movement of the Pacic plate with respect to the
Okhotsk plate is in the range of 8090 mm/year whereby the Pacic
plate moves under the Okhotsk plate. The accumulated strain at
plate boundary is occasionally released, resulting in large subduc-
tion earthquakes. In the off-shore Miyagi Prefecture (where the
rupture was initiated), many large earthquakes have occurred with
average recurrence of about 30 years and have caused severe
Contents lists available at ScienceDirect
journal homepage: www.elsevier.com/locate/soildyn
Soil Dynamics and Earthquake Engineering
0267-7261/$ - see front matter Crown Copyright & 2011 Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.soildyn.2011.06.006
n
Corresponding author.
E-mail address: subhamoy.bhattacharya@gmail.com (S. Bhattacharya).
Soil Dynamics and Earthquake Engineering 31 (2011) 16181628
destruction (e.g. 1978 M
w
7.7 earthquake). One distinct difference of
the 2011 Tohoku earthquake from previous ones is that the rupture
zone was signicantly larger and involved multiple nearby rupture
zones, such as off-shore Fukushima Prefecture and off-shore Ibaraki
Prefecture to south, and off-shore Iwate Prefecture to north (which
were usually thought to rupture individually, rather than simulta-
neously). In other words, the 2011 Tohoku earthquake resulted from
coupled co-seismic rupture of the above-mentioned major fault
segments; such simultaneous ruptures have not happened since the
869 Jogan earthquake [16], which caused massive tsunami along the
coastal areas of the Tohoku region.
The entire rupture process involved a large fault plane with
400500 km in length by 100200 km in width [10,19]. The fault
rupture models estimated by Shao et al. [19] (UCSB model) and by
the Geospatial Information Authority of Japan (2011) (GSI model)
are illustrated in Fig. 1 (note: many other fault plane models are
available). These models were derived by focusing on macro
features of observed ground deformation and long-period ground
motions. In particular, the GSI model approximates the overall
rupture process as a sequence of two sub-processes, initiated in the
northern plane and propagated to the southern plane. Although
the overall extent of the estimated fault rupture area differs
GSIs plane 1
M
w
= 8.8
GSIs plane 2
M
w
= 8.3
UCSBs plane
M
w
= 9.1
34
o
N
36
o
N
38
o
N
40
o
N
42
o
N
138
o
E 140
o
E 142
o
E 144
o
E
0 100 200
km
Kanto area
FKS013
IBR007
MYG008
Hypocentre
500 cm/s
2
500 cm/s
2
200 s
NS component
EW component
MYG008
Rrup = 65.5 km
VS30 = 273.6 m/s
FKS013
Rrup = 75.6 km
VS30 = 292.4 m/s
200 s
NS component
EW component
NS component
EW component
200 s
500 cm/s
2
500 cm/s
2
500 cm/s
2
500 cm/s
2
IBR007
Rrup = 51.2 km
VS30 = 292.8 m/s
Strong ground
motion generation
zone by Irikura &
Kurahashi (2011)
Fig. 1. Map of the Tohoku and Kanto regions and recorded ground motion data at three K-NET stations.
Haneda Airport
Kinshicho
KNG002 (K-NET)
CHB008 (K-NET)
Urayasu
CHB014 (K-NET)
Higashisuna
Tokyo Disneyland
CHB009 (K-NET)
Chiba
Shinkiba
Fig. 2. Locations of survey around the Tokyo Bay area (photo courtesy Google Maps).
S. Bhattacharya et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 16181628 1619
signicantly, depending on the analysis methods and the considered
frequency ranges for earthquake source inversion, locations for large
slips are generally similar among different approaches (i.e. an area
around the GSIs rupture plane 1 in Fig. 1). A signicant amount of
permanent ground deformation of up to about 30 m to the east [10]
occurred along the trench line off-shore of the Tohoku region (i.e.
eastern boundary of the GSI/UCSB fault planes in Fig. 1) and this was
mainly responsible for the generation of massive tsunami.
The macro-rupture process based on preliminary analyses
conducted by various research groups generally indicates that
the rupture initiated at the earthquake epicentre and propagated
southwards, rupturing the segments off-shore Fukishima and
Ibaraki Prefectures [10]. Moreover, a closer investigation into
strong ground motion generation in the short-period range by
Irikura and Kurahashi [12] suggests that the initiated rupture also
propagated towards deeper segments along the dip and triggered
several small patches of fault rupture with high stress drops in
areas between the UCSBs western boundary and the GSIs
western boundary shown in Fig. 1. (Note: high stress drops
usually produce intense short-period ground motions, and are
typical characteristics of deep inslab earthquakes.) In particular,
the source model by Irikura and Kurahashi [12] includes four
small rupture segments (see grey patches shown in Fig. 1): two
segments are placed in the west of the GSIs northern plane (i.e.
off-shore Miyagi Prefecture), while the other two segments are
placed in the west of the GSIs southern plane (i.e. one is off-shore
Fukushima Prefecture and the other is off-shore Ibaraki Prefec-
ture). These four segments are closer to the coastal areas of the
Tohoku region and are at deeper locations; generated seismic
waves travel through a so-called high-Q region [15], experiencing
less attenuation of seismic waves over distance, and thus results
in larger ground motions. Hence, the rupture and wave propaga-
tion process of the Tohoku earthquake was rather complex,
because multiple sub-ruptures were responsible for different
episodes of the overall rupture process (shallow versus deep)
and their effects on generated ground motions differ, depending
on the frequency range of interest (high versus low). Another
important consideration in understanding a relationship between
complex rupture process and observed ground motions is the
directivity effect. Observed ground motion time-series data tend
to be greater in peak and more concentrated in duration if the
rupture propagation direction coincides with the orientation from
the rupture source to the site, whereas they tend to be smaller in
peak and stretched over longer duration if the rupture direction is
opposite of the source-to-site orientation.
The complex rupture process described above can be illustrated
by inspecting recorded acceleration time-series data (both NS and
EW components) at three K-NET stations MYG008, FKS013 and
IBR007 (see Fig. 1). At the MYG008 station, two clear phases of
seismic wave arrivals can be seen, whereas at the FKS013 and
IBR007 stations, a single phase of seismic wave arrivals is featured.
In particular, for the IBR007 station, the acceleration amplitude or
strong motion part is more concentrated than the other two
stations. This is due to the combined effects of the complex rupture
process with multiple strong motion generation sources and the
source-to-site directivity; the rupture is initiated at the epicentre
and triggering of rupture moved towards south and west. By
considering a typical rupture propagation speed (which is about
80% of the seismic wave propagation speed at deeper locations),
arrivals of the triggered seismic waves at different parts of the fault
tend to be more concentrated in time, if a recording station is
located towards the direction of fault rupture process (e.g. IBR007),
whereas they become more spread out if a recording station is in the
opposite of the rupture propagation direction (e.g. MYG008). It is
also noted that the observed values of peak ground acceleration
(PGA) at these K-NET stations are large, exceeding the PGA of
1000 cm/s
2
(at some K-NET stations, MYG004 and MYG012,
recorded PGAs exceeded 2500 cm/s
2
). Large observed PGAs at
several stations in Miyagi, Fukushima and Ibaraki Prefectures are
partly due to proximity to localised strong ground motion genera-
tion areas along the coastline [12].
3. Field survey around Tokyo Bay area
Immediately following the earthquake, a eld investigation was
carried out along the coastline of Tokyo Metropolitan Government
and Chiba Prefecture. The survey area was limited to Urayasu,
ShinKiba, Chiba, Kinshicho, Haneda airport, Tokyo Disneyland and
Higashisuna along the river Naka (see Fig. 2). No apparent failures or
collapses of large superstructures were observed. The damage was
dominated by soil liquefaction. Surface liquefaction was observed at
many locations. The streets and car parks were lled up with a solid
mixture of brown and/or grey colour ne particle materials and
water. Clearly these materials erupted from underneath the ground
Fig. 3. Map of the Kanto region and recorded ground motion data at four K-NET stations.
S. Bhattacharya et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 16181628 1620
and were mostly ll materials and/or young alluvium. In the
following, observed ground motions and eld investigations of
liquefaction and ground settlement around the Tokyo Bay area are
summarised.
3.1. Observed ground motions around the Tokyo Bay area
Typical recorded ground acceleration time-series data at four
K-NET stations in the Tokyo Bay area are presented in Fig. 3. The
selected stations are at soft soils (typically, average shear wave
velocity in the upper 30 m V
S30
is less than 200 m/s, corresponding
to the NEHRP site class D/E boundary or E). These sites are susceptible
to soil liquefaction. The recorded ground motions at the four stations
are similar and have a PGA of about 100200 cm/s
2
. Note that the
duration of the strong ground motion part (or signicant cyclic load
reversals) is rather long, because of the large size of the event. These
ground motion characteristics (i.e. soft soil, relatively large ground
motion amplitude and long duration) are the necessary conditions for
liquefaction triggering [18,22]. Moreover, to inspect the frequency
content of the observed ground motion records and elastic seismic
demand characteristics, power spectral densities of two horizontal
components of the four records are shown in Fig. 4(a)(d) Average
5%-damped response spectra of the four records are shown in
Fig. 4(e). The dominant period ranges of the recorded ground motions
at sites around the Tokyo Bay area were around 0.751.25 s, and
high-frequency content of these records was relatively low.
3.2. Observed liquefaction in parks and streets
Fig. 5 shows the condition of the 10-lane road to Tokyo Disney-
land and parking as noted on the 12th March morning before the
repair work commenced. At one location the road surface buckled,
Fig. 4. Power spectral density of ground motion data (ad) and 5%-damped acceleration response spectra of ground motion data (e) at four K-NET stations in the Tokyo Bay area.
S. Bhattacharya et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 16181628 1621
and other surface liqueed soil can be seen at the surface, while in
the parking area widespread liquefaction was observed. Fig. 6 shows
the brittle failure of a thin black top asphalt slab of the road to
Tokyo Disneyland. Clearly a compressive force was developed on the
earth surface causing the weak pavement slab to buckle upwards.
The possible causes may be due to a combination of vertical up-
down motion, dynamic surface wave propagation effects or accu-
mulated permanent compressive strains in the surface soils. These
observations of buckled pavements were not very common in the
damaged area surveyed (i.e. localised), which may hint that these
zones were those where the construction of the road or compaction
of the ground was less than perfect.
Widespread liquefaction was observed in certain areas with
plenty of evidence of sand boils. Fig. 7 shows photographs of sand
boils on the morning of 12th March. While Fig. 7(left) shows the
sand boils in the Takasu Park, Fig. 7(right) shows the liquefaction
in the paved car park of the Disneyland amusement park. The
liqueed and ejected soil consisted of different types of materials
ranging from pure sand (brown colour) with small nes content
to grey silty sand, and also in some locations dredged recycled
material. The type of ejected materials varied from location to
location and seemed to depend on the dates when the land
reclamation or ground improvement was carried out. The boiled
material also differed from place to place and was highly
dependant on the specic gravity of the solid grains constituting
the soil. Liquefaction caused settlement, and the amount varied
from location to location, depending on the type of ground
improvement carried out in the locations. Generally, the parks
incurred liquefaction disturbance more than the built up areas.
More discussion of ground improvement practices is given later in
the paper. Fig. 8 shows the enormous quantity of ejected sand in
the main roads. It mainly erupted from the edges where the road
met the footpath or the cycle path. It must also be mentioned that
as the duration of the earthquake was rather long (approximately
100200 s, see Fig. 3), the soil was subjected to a signicant
number of load reversals beyond that normally expected. This
must have contributed to the unusually high amounts of liqueed
material ejected to the ground surface. Fig. 8 shows the repair
work being carried out on one of the main roads where a vast
amount of ejected liqueable material was on the road surface.
Fig. 5. Black top exible pavement to Tokyo Disneyland.
Fig. 6. Buckled pavements in the road to Tokyo Disneyland.
Fig. 7. Observed liquefaction in the park (left) and parking area of Disneyland (right).
S. Bhattacharya et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 16181628 1622
For illustration, a simplied stress-based liquefaction trigger-
ing analysis [18,22] was carried out by considering a typical soil
prole in Urayasu, where widespread liquefaction and sand boils
were observed. Specically, a probabilistic model of liquefaction
initiation, developed by Cetin et al. [7], was adopted. The results
are shown in Fig. 9. The calculated probability of liquefaction over
depth up to 20 m is presented in the third sub-gure in Fig. 9.
Three analysis cases were conducted by considering the different
moment magnitude values, because the developed model has not
been validated/calibrated against a mega thrust M
w
9.0 subduc-
tion event. The analysis results indicate that soil liquefaction is
highly likely to be triggered at shallow depths (up to 6 m) and
medium depths (1216 m), and the effects of the moment
magnitude are not so signicant at these depths. The main reason
for the high liquefaction potential for this prole is the low N
count (less than 5). It is noted that the liquefaction triggering is
not necessarily a reliable indicator of the extent of soil liquefac-
tion and settlement; more detailed analysis is needed (which is
beyond the scope of this study).
3.3. Liquefaction induced damage to light structures, walkways,
trafc signal posts and sign posts
In some locations, liquefaction caused damage to light struc-
tures, such as trafc signal posts, lamp posts and electrical poles,
foot paths, walkways and plinth protection structures of buildings.
Gaps were formed between sidewalks of buildings and the building
itself, causing noticeable relative settlement. Fig. 10 shows a
photograph illustrating the settlement of the footpath in a multi-
storied building in Urayasu. It was observed that the relative
settlement between the footpath and the building was variable
due to the different types of foundations. In some buildings minor
tilting associated with differential settlement was observed. Minor
tilting also limited the damage to the superstructure. Fig. 11 shows
some light weight structures where signicant tilt can be noted.
Many of these light weight structures were supported on small
footings connected through grade beams. The relative settlement
varied between 5 and 15 cm.
3.4. Liquefaction induced damage to yovers and elevated highways
Many of the elevated highways, yovers and bridges were
surveyed. Damage to superstructures or piers of these structures
was not observed, which can be considered as a success in
comparison with the performance of Hanshin expressway during
the 1995 Kobe earthquake. However, it should be noted that
ground motion amplitudes in the Tokyo area were much lower
than those observed in Kobe, but the duration of the strong
motion in 2011 Tohoku earthquake was signicantly longer than
the Kobe motion. The downtime was minimal and these struc-
tures became operational within a few hours after inspection.
This allowed the survey team to travel by train to the site almost
Fig. 8. Repair works of the road along with ejected material at the surface of the road.
D
e
p
t
h

(
m
)
0
2
4
6
8
10
12
14
16
18
20
Liquefaction
probability
0.0 0.5 1.0
N count
Soil type
0 10 20 30
Max M
w
= 8.5
Max M
w
= 8.0
Max M
w
= 9.0
Sandy silt
Sandy silt
Fine sand
Sandy silt
Silty fine sand
Input parameter:
PGA: 142.9 cm/s
2
V
S12
: 123 m/s
Water table: 1 m
Dry density: 1.76 g/cm
3
Wet Density: 1.92 g/cm
3
Fig. 9. Typical soil prole in Urayasu (for location see Fig. 2).
S. Bhattacharya et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 16181628 1623
immediately. Fig. 12 shows the liquefaction induced damage
around the foundations of the piers for the elevated yovers in
Maihama. The damage was due to uneven settlement of the
surrounding ground relative to the pier; the relative settlement
varied between 10 and 30 cm. The foundations of these elevated
highways were piles.
3.5. Uplift of manholes
As was observed in previous earthquakes (such as 2004
Niigata Ken-Chuetsu earthquake), many manholes popped out
of the ground, causing damage to the water and sewerage
pipelines. Fig. 13 shows a photograph of an uplifted manhole. A
plausible reason is a combination of the settlement of the
surrounding ground and the upward buoyancy force generated
by liquefaction.
4. Discussion
The failures (structural and/or geotechnical) in past earth-
quakes have shown the shortcomings of design methodologies
and construction practices. Post earthquake reconnaissance inves-
tigations have led to the identication of limitations of engineer-
ing analysis, design and construction practices, as outlined in
Table 1. This section of the paper summarises the key observa-
tions from the present eld investigations and discusses the
lessons learnt. The observations are also compared and contrasted
with the previous earthquakes.
Fig. 11. Differential settlement of light structures: (a) Tilting of vending machines due to subsurface liquefaction; (b) tilting of signal posts; (c) severe differential
settlement of a transformer box near Maihama Station.
Fig. 12. Damage around the footing of the elevated highways.
Fig. 10. (a) Gap forming between a building and a sidewalk of a building; (b) local settlement depression of a hotel; (c) settlement of surrounding ground next to a
bridge pier.
S. Bhattacharya et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 16181628 1624
4.1. Ground response
Fig. 14 compares ground motion time-series data at two
recording sites: CHB008 and CHBH14 (see Figs. 2 and 3 for
locations). CHBH14 station is a part of the KiK-NET, and both
the surface and borehole data (recording at the bedrock) are
available. Fig. 14(c) shows the bedrock motion recorded at the
CHBH14 site, which can be considered to be the same as the site
CHB008 for all practical analysis purposes. Fig. 14(a) and (b), on
the other hand, shows the effect of wave propagation on soil
layering at two locations. The comparison of surface-to-borehole
ground motion data at these locations demonstrates the soil
amplication in the top soft soil layers. At these locations, the
ground motion was amplied by more than 4 times. Fig. 14(a) and
(b) also demonstrates the long-duration cyclic loading with rela-
tively large amplitudes at CHB008 and CHBH14. Both amplitudes
and large number of load reversals contribute signicantly to the
liquefaction triggering [22].
4.2. Effects of ground improvement
It was clear from the survey that the severity of liquefaction
varied enormously at different locations. There was a construc-
tion site in Hanamigawa-Ku (Chiba) with a stark contrast. The
part of the site where the ground was improved did not liquefy at
all, while the remaining part liqueed. Sand compaction piles
(SCP) were used at the site for ground improvement. Another
contrasting example was the excellent performance of Haneda
airport, which was built on reclaimed land using various ground
improvement techniques. There was practically no downtime in
Haneda airport, and ights were operational from the very next
day of the earthquake.
Table 1
Historical development of earthquake engineering practice.
Earthquake Remarks Post earthquake developments
1908 Reggio Messina
earthquake (Italy)
120,000 fatalities. A committee of nine practising engineers
and ve professors were appointed by Italian government to
study the failures and to set design guidelines.
Base shear equation evolved i.e. the lateral force exerted on
the structure is some percentage of the dead weight of the
structure, (typically 515%).
1923 Kanto earthquake
(Japan)
Destruction of bridges, buildings. Foundations settled, tilted
and moved.
Seismic coefcient method (equivalent static force method
using a seismic coefcient of 0.10.3) was rst
incorporated in design of highway bridges in Japan (MI
1927).
1933 Long Beach
earthquake (USA)
Destruction of buildings specially school buildings. UBC (1927) revised. This is the rst earthquake for which
acceleration records were obtained from the recently
developed strong motion accelerograph.
1964 Niigata earthquake
(Japan)
Soil can also be a major contributor of damage. Soil liquefaction studies started.
1971 San Fernando
earthquake (USA)
Bridges collapsed, dams failed causing ood. More soil effects
were observed.
Liquefaction studies intensied. Bridge retrot studies
started.
1994 Northridge
earthquake (USA)
Steel connections failed in bridges. Importance of ductility in construction realised. Signicant
damage potential due to near-fault motions was
recognised.
1995 Kobe earthquake
(Japan)
Massive foundation failure. Soil effects were the main cause
of failure.
Downward movement of a slope (lateral spreading) is said
to be one of the main causes of foundation failure. JRA
(1996) code modied (based on lateral spreading
mechanism) for design of bridge piles.
1999 Chi-Chi earthquake
(Taiwan)
Many bridges collapsed as they were located close to the
faults.
Special care must be undertaken while designing
important structure in the vicinity of the plate boundaries
and faults
1999 Koceli earthquake
(Turkey)
Damage to Bolu tunnel due to fault movement. Damage to
buildings and bridges. However, buildings conforming the
design codes performed well.
Research on fault induced failures intensied.
2001 Bhuj earthquake
(India)
Large-scale destruction. Good performance of some new
jetties of the Kandla port. Tilting of the Kandla Tower
building without any damage.
Large diameter piles concrete performed better than small
diameter piles. Steel piles lled in concrete also performed
better.
2004 Sumatra earthquake
and tsunami
Destruction to built environment due to earthquake and
giant tsunami waves.
New research focused on tsunami warning systems.
Fig. 13. Uplift of manholes.
S. Bhattacharya et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 16181628 1625
Various methods are available for ground improvement in Japan,
such as the sand drain method and the sand compaction pile method.
4.2.1. Sand drain method
In this method, vertical drains either of coarse soil or permeable
synthetic variety are installed through liqueable soils to alleviate
the liquefaction problems. The principle is to speed the dissipation
of the excess pore water pressure before reaching full liquefaction.
This method was used in the Haneda airport underneath some of
the runways. A good discussion on the vertical drains and the
associated methodology can be found in Yasuda et al. [21], Orense
et al. [17], Adalier et al. [2] and Brennan and Madabhushi [6].
4.2.2. Sand compaction pile (SCP)
This method was originally developed in Japan in the 1960s to
treat liqueable soil. It essentially induces densication and sub-
sequent compaction of the soil matrix. The equipment used is
similar to that used for sand drain installation, where large amounts
of sand are injected into the ground through a combination of
repeated driving down and extracting motions of large diameter
steel pipes. Essentially, when the equipment reaches a particular
chosen depth, the equipment is withdrawn and the sand is poured
for a predetermined depth through its mandrel. Subsequently, using
the vibrator attached to the top of the mandrel, sand is compacted
and allowed to expand across the diameter. The injection of sand
increases the relative density of the soil and thus strengthens the
susceptibility against liquefaction. Following the ground improve-
ment, proper foundations (pile or mat) are put in place. For the
ongoing construction site in Hanamigawa-Ku referred to earlier, the
sand compaction piles were 700 mm in diameter and extended to a
depth of 1620 m. Sand compaction piles were used to treat the
Haneda airport. The latter experienced the earthquake without
suffering signicant liquefaction damage. A good discussion of the
methodology of SCP can be found in Akiyoshi et al. [1].
4.3. Good performance of the Haneda airport using the various
methods of ground improvement
The airport was constructed on reclaimed land from the sea
using various methods of ground improvement: sand compaction
pile, sand drain, deep cement mixing and soil replacement.
Ground improvement was one of the major construction hurdles
in this project. Out of the total 36 months construction period, 12
months were devoted to ground improvement. A good discussion
of the methods of ground improvement employed in the Haneda
airport can be found in JGS [14].
4.4. Choice of recycled material for land reclamation
Land reclamation is very common in Japan, as 70% of Japan is
covered by mountains. Therefore, the choice and procurement of
Fig. 14. Amplication of ground motion as evidenced from the recordings at two stations; (a) ground motion as recorded at the surface at CHB008 location (K-NET);
(b) ground motion as recorded at the surface at CHBH14 location (KiK-NET); (c) bedrock motion recorded at the CHBH14 site.
S. Bhattacharya et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 16181628 1626
good ll materials are very important. The ll materials used at
sites that performed very well during the earthquake consisted of
sands, stones and light weight cement treated dredged soils with
and without air. It has been estimated that 6 million m
3
of cement
treated dredged soils were used during the construction of
Haneda airport. They are considered to be very light and strong
recycled materials. Cement treated dredged soil without air is
heavier than water but lighter than ordinary soils. On the other
hand, cement treated dredged soil with air, known as Super Geo
Material (SGM), is very light and almost the same weight as
water. These soils were used as land reclamation materials behind
the sea wall, underneath the runway and taxiway. The earthquake
showed that these soils are very effective in reducing consolida-
tion/settlement and avoiding ground failure of sea walls.
4.5. A case study: good performance of piled-supported 15-storey
building
Based on the preliminary investigation and experience from
the past earthquakes, it was apparent that the soil remained
liqueed for some time. A survey was carried out in a modern
(2005) 15-storey condominium (mansion) in Higashisuna in the
Sunamachi area on the banks of Naka River (Nakagawa). The
condominium has 236 apartments and runs parallel to the Tokyo
Metropolitan Area highway. There was no damage to the super-
structure, see Fig. 15 for photograph taken after the earthquake
and the plan of the building. It may be noted that the building is L
shaped but there is no structural connection between the two
sections.
The building is supported on concrete piles having 12 m
diameter and is 5354 m long. The diameter of the pile at the base
ranges from 1.4 to 3.5 m to have an enhanced end-bearing.
The load carrying capacity of the pile ranges between 2356 and
15,009 kN. Each column was supported on one of these large
diameter piles. The superstructure is a steel frame encased within
reinforced concrete and the interior is light wood partition wall.
The long side of the building runs along the river Naka (Naka-
gawa). Lateral spreading was not observed along the river banks
due to good performance of river bank protection work.
Table 2 lists the diameter of piles of ten pile-supported
structures in liqueable soils and their observed performance
following a strong earthquake. The case records were taken from
the performance of structures following the 1964 Niigata earth-
quake, the 1995 Kobe earthquake and the 2001 Bhuj earthquake.
It may be noted that the 1960s design concept of few small
diameter piles (typically 0.30.6 m) were gradually replaced by a
few large diameter piles in the 1990s. The poor performance in
this context signies formation of plastic hinges in the pile and
cracking. Further details of these case studies and failure modes
can be found in Dash et al. [8,9], Bhattacharya and Madabhushi
[3] and Bhattacharya et al. [4,5]. In most of these cases it has been
estimated that soil liqueed up to 710 m of depth. It is quite
clear from the table that small diameter piles performed poorly.
Bhattacharya et al. [4] carried out a comprehensive study of 15
case histories of piled foundation performance based on buckling
analysis parameters. Essentially, the study showed that a pile is
laterally unsupported in the liqueable zone due to the removal
of lateral connement by the soil owing to liquefaction. The
slenderness ratio, L
eff
/r
min
, of the pile in the region that could
become unsupported is used to classify pile foundation perfor-
mance. L
eff
is the Eulers effective length of an equivalent pin
ended strut and r
min
is the minimum radius of gyration of the pile
section. The study showed that a line representing a slenderness
ratio (L
eff
/r
min
) of 50 could distinguish between poor and good pile
performance. This line is of some signicance in structural
Fig. 15. Pile-supported condominium in Sunamachi area (photo and plan view).
Table 2
Case histories of pile foundation performance during past earthquakes.
Case history and reference Pile diameter Performance
(good/poor)
NFCH pile-supported building during 1964 Niigata earthquake, see [11] 0.35 m dia RCC hollow Poor
Pile-supported Yachiyo Bridge during 1964 Niigata earthquake, see [11] 0.3 m dia RCC Poor
Pile-supported N.H.K building during 1964 Niigata earthquake [11] 0.35 m dia RCC Poor
Pile-supported Showa Bridge during the 1964 Niigata earthquake [5] 0.6 m dia Steel Poor
Pile-supported LPG tank 101 during 1995 Kobe earthquake [13] 1.1 m dia RCC Good
Pile-supported LPG tank 106 and 107 during 1995 Kobe earthquake [13] 0.3 m dia RCC hollow Poor
Pile-supported 14 storey building during 1995 Kobe earthquake [20] 2.5 m dia RCC Good
Pile-supported Kandla Port Tower during 2001 Bhuj earthquake [8] 0.4 m dia RCC Poor
Pile-supported old jetty in Kandla port during the 2001 Bhuj earthquake [3] 0.5 m dia RCC Poor
Pile-supported new jetty in Kandla port during the 2001 Bhuj earthquake [3] 1.0 m RCC lled steel tube Good
S. Bhattacharya et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 16181628 1627
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 modied by
induced bending moments. This slenderness ratio of 50 signies
a length to diameter of about 12 for RC columns. Furthermore, the
study carried out by Dash et al. [9] showed that large diameter piles
were safe against the bending, bucking and bendingbuckling
interaction types of failure. The good performance of the building
described in Fig. 15 further reinforces the hypothesis that few large
diameter piles are better that many small diameter piles.
5. Conclusions
The 11th March 2011 moment magnitude M
w
9.0 off the coast
Tohoku earthquake was the largest earthquake ever recorded in
Japan. It was an interface subduction earthquake with a low-angle
reverse mechanism, and was caused by tectonic movements of the
North American (Okhotsk) plate and the Pacic plate. Intense
tremors were felt in the Kanto (Tokyo) area, which is about 350
400 km from the epicentre. Analysis of the recorded strong motion
data showed that the dominant period ranges of the sites around the
Tokyo Bay area are around 0.751.25 s. The high-frequency content
of these records was relatively low. The duration of the earthquake
was unusually long at approximately 100200 s.
Widespread liquefaction was observed in the Tokyo bay area
especially in the zones of reclaimed land, ll areas or sites having
young alluvium. Liquefaction caused ground failures and also
damage to the built environment. Typical examples of damage
included tilting of light structures, such as trafc signal posts,
while gaps formed between sidewalks of buildings and the
buildings. The long-duration effects of ground motion contributed
signicantly to the liquefaction induced damage. Liquefaction
susceptibility and the damage caused by liquefaction seemed to
be dependant on the age of ll (i.e. when the land reclamation
was carried out), type of ll material (source whether it is
dredged material, pure sand or silty sand) and the type of ground
improvement carried out.
Acknowledgements
The rst author would like to acknowledge the support
extended by Professor Kohji Tokimatsu (Centre for Urban Earth-
quake Engineering, Tokyo Institute of Technology) and Engineer-
ing and Physical Sciences Research Council (EPSRC) for support
through the Grant EP/H015345. The data available through the
K-NET database and NEIC data catalogue is also acknowledged.
The authors also thank Dr Nick Alexander for fruitful discussions.
The third author is grateful to Professor Kojiro Irikura for his
thorough explanations on the rupture process of the 2011 Tohoku
earthquake.
References
[1] Akiyoshi T, Fuchida K, Matsumoto H, Hyodo T, Fang HL. Liquefaction analyses
of sandy ground improved by sand compaction piles. Soil Dynamics and
Earthquake Engineering 1993;12:299307.
[2] Adalier K, Elgamel A, Meneses J, Baez JI. Stone columns as liquefaction
countermeasure in non-plastic silty soils. Soil Dynamics and Earthquake
Engineering 2003;23:57184.
[3] Bhattacharya S, Madabhushi SPG. A critical review of methods of pile design in
seismically liqueable soils. Bull of Earthquake Engineering 2008;6:40747.
[4] Bhattacharya S, Madabhushi SPG, Bolton MD. An alternative mechanism of
pile failure in liqueable deposits during earthquakes. Geotechnique
2004;54(April issue, 3):20313.
[5] Bhattacharya S, Bolton MD, Madabhushi SPG. A reconsideration of the safety
of the piled bridge foundations in liqueable soils. Soils and Foundations
2005;45(4):1326.
[6] Brennan AJ, Madabhushi SPG. Liquefaction remediation by vertical drains
with varying penetration depths. Soil Dynamics and Earthquake Engineering
2006;26:46975.
[7] Cetin KO, Seed RB, Der Kiureghian A, Tokimatsu K, Harder Jr. LF, Kayen RE, et al.
Standard Penetration Test-based probabilistic and deterministic assessment of
seismic soil liquefaction potential. Journal of Geotechnical and Geoenviron-
mental Engineering 2004;130:131440.
[8] Dash SR, Gobindraju L, Bhattacharya S. A case study of damages of the Kandla
Port and Customs Ofce tower supported on a matpile foundation in
liqueed soils under the 2001 Bhuj earthquake. Soil Dynamics and Earth-
quake Engineering 2008;29(2009):33346.
[9] Dash SR, Bhattacharya S, Blakeborough A. Bendingbuckling interaction as a
failure mechanism of piles in liqueable soils. Soil Dynamics and Earthquake
Engineering 2010;30:329.
[10] Geospatial Information Authority of Japan (GSI). Crustal deformation and
fault model obtained from GEONET data analysis; 2011. [Available at:
/http://www.gsi.go.jp/cais/topic110313-index-e.htmlS].
[11] Hamada M. Large ground deformations and their effects on lifelines: 1964
Niigata earthquake. Case Studies of liquefaction and lifelines performance
during past earthquake. Technical Report NCEER-92-0001, Volume-1, Japa-
nese case studies, National Centre for Earthquake Engineering Research,
Buffalo, NY; 1992.
[12] Irikura K, Kurahashi S. Source model for generating strong ground motions
during the 11 March 2011 off Tohoku, Japan earthquake. In: Proceedings of the
Japan geoscience union international symposium. Makuhari, Chiba, Japan; 2011.
[13] Ishihara K. Terzaghi oration: Geotechnical aspects of the 1995 Kobe earth-
quake. In: Proceedings of the ICSMFE, Hamburg; 1997, p. 204773.
[14] Japanese Geotechnical Society. Geotechnical engineering of Tokyo interna-
tional airport offshore expansion project-slide library, 1999.
[15] Kanno T, Narita A, Morikawa N, Fujiwara H, Fukushima Y. A new attenuation
relation for strong ground motion in Japan based on recorded data. Bulletin
of the Seismological Society of America 2006;96:87997.
[16] Minoura K, Imamura F, Sugawara D, Kono Y, Iwashita T. The 869 Jogan
tsunami deposit and recurrence interval of large-scale tsunami on the Pacic
coast of northeast Japan. Journal of Natural Disaster Science 2001;23:838.
[17] Orense RP, Morimoto I, Yamamoto Y, Yumiyama T, Yamamoto H, Sugawara K.
Study on wall-type gravel drains as liquefaction countermeasure for underground
structures. Soil Dynamics and Earthquake Engineering 2003;23(1):1939.
[18] Seed HB, Idriss IM. Simplied procedure for evaluating soil liquefaction potential.
Journal of the Soil Mechanics and Foundations Division 1971;97:124973.
[19] Shao G, Li X, Ji C, Maeda T. Preliminary Results of the March 11, 2011 M
w
9.1 Honshu Earthquake. University of California, Santa Barbara; 2011.
[Available at: /http://www.geol.ucsb.edu/faculty/ji/big_earthquakes/2011/
03/0311/Honshu_main.htmlS].
[20] Tokimatsu K, Mizuno H, Kakurai M. Building damage associated with
geotechnical problems. Special issue of Soils and Foundations, Japanese
Geotechnical Society; 1996, p. 21934.
[21] Yasuda S, Ishihara K, Harada K, Shinkawa N. Effect of soil improvement on
ground subsidence due to liquefaction. Soils and Foundation 1996:99107.
[22] Youd TL, Idriss IM, Andrus RD, Arango I, Castro G, Christian JT, et al.
Liquefaction resistance of soils: summary report from the 1996 NCEER and
1998 NCEER/NSF workshops on evaluation of liquefaction resistance of soils.
Journal of Geotechnical and Geoenvironmental Engineering 2001;127:81733.
S. Bhattacharya et al. / Soil Dynamics and Earthquake Engineering 31 (2011) 16181628 1628

You might also like