You are on page 1of 15

Dong, Y. P. et al. (2016). Gotechnique 66, No. 1, 115 [http://dx.doi.org/10.1680/jgeot.14.P.

234]

Finite-element analysis of a deep excavation case history


Y. P. DONG , H. J. BURD and G. T. HOULSBY

The design of deep excavations requires careful consideration of the influence of various soil/structure
interaction mechanisms and detailed issues relating to the construction processes and the mechanics of
the soil. Finite-element analysis provides a useful design tool for deep excavations, but care needs to
be taken to ensure that an appropriate level of detail is included in the model. This paper describes a
three-dimensional finite-element analysis of a deep excavation supported by a diaphragm wall, recently
constructed in Shanghai. The principal purpose of the study is to investigate the level of detail that is
required in the finite-element model to obtain results that provide a realistic representation of the wall
and ground movements measured during the construction process. Studies are conducted on (a) the
influence of the soil constitutive model on the quality of the results; (b) procedures to model the effect of
post-cure shrinkage in the concrete floor slabs; (c) procedures to model the construction joints in the
diaphragm wall; (d ) the relative merits of using shell and solid elements to model the diaphragm wall;
and (e) the sensitivity of the analysis to the assumed initial horizontal stresses in the soil.
KEYWORDS: case history; excavation; finite-element modelling; retaining walls; soil/structure interaction

available software and hardware. Experience gained during


calibration exercises, in which the results of finite-element
analyses are compared with field data, may be used to suggest
appropriate numerical procedures to adopt and pitfalls to
avoid.
To obtain a satisfactory numerical model of the performance of a deep excavation, a detailed three-dimensional
(3D) model is typically required (Gourvenec et al., 2002;
Zdravkovic et al., 2005; Lee et al., 2011). Furthermore, the
analysis needs to take account of the small strain nonlinearity of soil (Simpson, 1992; Potts & Zdravkovic, 2001),
potential post-cure thermal effects associated with the floor
slabs that act to support the retaining structures (Whittle
et al., 1993), and the initial stress state in the ground (Potts
& Fourie, 1984). Other issues to be considered include the
choice of element type (i.e. continuum or shell) to model the
retaining wall, and the development of an appropriate approach to model the structural influence of any construction
joints in the retaining wall (Zdravkovic et al., 2005).
This paper describes a detailed analysis of a complex
deep excavation case history (the basement excavation for
Shanghai Xingye Bank building) using Abaqus V611. This
project involved the top-down construction of a deep
excavation, supported by a diaphragm wall. The purpose of
the current study is to investigate the influence of various
modelling approaches and procedures on the computed
behaviour. Studies are conducted on the relative merits of
alternative approaches for modelling the soil, the retaining
wall and the supporting structures. Detailed field measurements are available for this project (Xu, 2007); these data are
used to assess the reliability of the finite-element results.

INTRODUCTION
The design of deep excavations requires careful consideration
of the strength and stability of the various structural elements
at all stages during the construction process. In addition, the
ground movements induced by the excavation need to be
carefully controlled, to ensure that damage to any nearby
buildings and services is kept within acceptable levels. The
performance of a deep excavation depends on the method of
construction as well as the local ground conditions. Making
reliable predictions of performance often presents a considerable challenge.
A substantial body of field data from previous deep excavation projects is available in the literature, for example,
in the UK (Skempton & Ward, 1952; Wood & Perrin, 1984;
Simpson, 1992), the USA (Finno & Nerby, 1989; Finno
et al., 1989; Finno & Bryson, 2002), and Shanghai, China
(Liu et al., 2005, 2011; Wang et al., 2005; Xu, 2007; Ng et al.,
2012; Tan & Wei, 2012). Case histories of this sort provide
valuable information on the performance of various forms of
retaining system that can be used to calibrate finite-element
modelling procedures; information of this sort may also be
used to establish an appropriate level of confidence in the
results of finite-element analysis when used as part of the
design process for deep excavations.
Rapid recent advances in computing resources open up
new possibilities for the use of finite-element modelling for
the routine design of deep excavations. Considerable care
needs to be taken, however, to ensure that appropriate procedures are employed. If the model is too simplistic then
the results will be unreliable. Alternatively, if an attempt is
made to develop a model with an excessive level of detail,
then difficulties may arise in the selection of material and
construction parameters, or in limitations imposed by the

CASE HISTORY DESCRIPTION


General description
The Shanghai Xingye Bank is a high-rise building (825 m
high) with a three-level deep basement. The structure employs a reinforced concrete frame, founded on bored piles
(Xu, 2007). The basement excavation is approximately
80 m  90 m in plan (Fig. 1). The excavation depth, as
shown in Fig. 2, is 142 m on the west side and 122 m on
the east side. The excavation is adjacent to 15 densely packed

Manuscript received 19 November 2014; revised manuscript


accepted 14 July 2015. Published online ahead of print 14
September 2015.
Discussion on this paper closes on 1 June 2016, for further details see
p. ii.
 SingaporeMIT Alliance for Research and Technology, Singapore;
former DPhil student at University of Oxford, UK.
Department of Engineering Science, University of Oxford,
Oxford, UK.

1
Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

DONG, BURD AND HOULSBY

Drainage pipeline
Telephone cable

(B7)

(B6)

Zhongnan
building

(B5)

Lainhe
building

Gas pipeline

(B4)

Water supply pipeline


Electric power pipeline

Soil cement columns

Hankou Road

East China Architecture


Design Institute

Diaphragm wall
144 m
124 m
Xingye Bank

Spiral injection piles


( 600 mm)

Soil cement columns


(62 m wide)

Communication Bank

(CB)

HSBC Bank

Middle Jiangxi Road

(ECADI)

Custom House
Middle Sichuan Road

Electric power pipeline

Sanjing Bank

(SJB)

(B3)

(B2)

(B1)
Xincheng building

Water supply pipeline


Sewerage pipeline
Telephone cable
Drainage pipeline

Fuzhou Road

Root piles ( 300 mm)

10 15 20 m

Fig. 1. Plan view of the deep excavation (Xu, 2007)

14050 m

9450 m

Crab

4950 m
Soil cement
columns

Spiral injection piles


0100 m

0200 m
1
2

Temporary
strut
1350 m

Steel lattice
column

Steel pipe
( 0609 m)

6650 m

7100 m

Temporary
strut
10700 m

1350 m
Temporary
strut

10400 m

Root piles
( 03 m)

3900 m

Bottom slab
12400 m
14400 m

Concrete cushion
(02 m thick)

Soil cement
columns

5 11

Soil cement columns

5 12

Diaphragm wall
(10 m thick)

Bored pile ( 08 m)Bored pile ( 09 m)


3

60 m deep below ground level

Fig. 2. Cross section AA, see Fig. 1 (Xu, 2007)

Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

Diaphragm wall
(08 m thick)

FINITE-ELEMENT ANALYSIS OF A DEEP EXCAVATION CASE HISTORY


buildings (eight of which have historic significance) and
several existing underground service pipes.
The retaining system, shown in Fig. 2, consists of a diaphragm wall (with thickness varying between 08 m and
10 m), vertical columns and piles (08 m and 09 m in diameter, 60 m deep), three levels of horizontal concrete floor
slabs (015 m thick), a grid of reinforced concrete beams
(05 m  08 m in section) and several temporary struts.
A plan view of the ground floor slab and the grid of supporting beams is shown in Fig. 3. Openings in the floor slab
were designed to facilitate the removal of the excavated soil
and to provide lighting and ventilation to the lower levels.
The excavation was constructed using a typical top-down
approach. The sequence is summarised in Table 1.
Soil modelling procedures
The Xingye Bank is located at a site in Shanghai which is
underlain by thick, relatively soft, quaternary alluvial and
marine deposits known as Shanghai Clay. As described later,
these deposits include various clay and silty clay layers with a

3
9

low coefficient of permeability (typically 10 m/s). The


linear dimensions of the excavation are relatively large (of the
order of tens of metres) and, although some dissipation of
excess pore pressures is likely to occur during the construction process, it is assumed in the current analysis that these
drainage effects are minimal. The analyses described in the
current paper are therefore based on the assumption of
undrained soil behaviour.
It is noted that the ground movements that are caused by
deep excavation construction in Shanghai Clay are typically
observed to vary with time. Liu et al. (2005), for example,
report field data relating to a 17 m deep diaphragm wallsupported excavation in Shanghai. These authors conclude
that the observed time-dependency in the measured ground
settlements around the excavation provide evidence that significant dissipation of excess pore pressures occurred during
the construction process. Conversely, Tan & Wei (2012)
suggest, in connection with a separate set of deep excavation
field data in Shanghai Clay, that the time-dependent nature
of the observed post-construction settlements is a consequence of the known tendency of soils in this region to

Reinforced concrete
strut

Reinforced concrete
strut

Steel strut

Fig. 3. Plan view of the ground floor slab and supporting beams (Xu, 2007)

Table 1. Construction sequence (Xu, 2007)


Stages

Period: dates

Interval:
days

02/03/2002
06/10/2002
07/10/2002
19/10/2002
20/10/2002
11/12/2002
12/12/2002
30/12/2002
31/12/2002
27/02/2003
28/02/2003
24/03/2003
25/03/2003
11/05/2003
12/05/2003
10/07/2003
11/07/2003
24/09/2003
25/09/2003
21/10/2003
22/10/2003
11/12/2003

218

2
3
4
5
6
7
8
9
10
11

Construction activities
Install diaphragm walls, pile foundations; conduct ground improvement and dewatering

53

Excavate to elevation 15 m first, then excavate to elevation 53 m with slope ratio 1:15; slope
shoulder 10 m on the west and south side, and 8 m on the other two sides
Cast the beams and slabs for the top level of the basement

19

Excavate the berms surrounding the wall remaining from the previous stage to elevation 53 m

59

Cast the 1st level beams and floor slabs and the ground floor slab

25

Excavate to elevation 855 m

48

Cast beams and slabs for the 2nd level, and structures for the first floor above ground level

60

Excavate to elevation 107 m first, then excavate to 124 m with slope ratio 1:15; excavate berms
to elevation 113 m
Cast the bottom slab (2 m thick) and add temporary struts for 3rd level; construct the structures
for the 2nd floor above ground level
Excavate the remaining soil to elevation 144 m (west side) and 124 m (east side) respectively

13

76
27
51

Cast the bottom slab on the west side; remove the temporary struts; construct the other structures
of the basement

Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

DONG, BURD AND HOULSBY

Geotechnical conditions and soil properties


According to the original site investigation report (SGIDI,
1997), the site is located on a flat coastal plain, with ground
elevation between 480 m to 387 m. The water table is
between 05 m and 1 m below the ground surface. The site is
underlain by deposits of Shanghai Clay. The geological
profile and soil properties from the site investigation report
are shown in Fig. 4. The soil profile is divided into nine layers
according to differences in soil characteristics, physical
and mechanical properties. The natural water content of
the clay and the silty clay layers is close to, or in some cases
higher than, the liquid limit, suggesting that the soil is
either normally consolidated or lightly overconsolidated.
The undrained shear strength, su, determined from field vane
shear testing, is significantly higher than the values normally
associated with clay at the liquid limit, suggesting that the
clay is likely to be sensitive.
The data in Fig. 4 are insufficient to calibrate a soil model
in which the small strain non-linearity is included. In addition, the data only provide information on the undrained
shear strength to a depth of about 24 m below the ground
level, but the numerical analysis requires strength data to
a greater depth. To supplement the information provided in
the original site investigation report, additional data were
collected from published soil properties on Shanghai Clay, as
described below.
A set of undrained shear strength data (Dassargues et al.,
1991) measured using shear box tests on soils from the
central zone of Shanghai, and additional undrained shear
strength data determined using the field vane at two separate
sites in Shanghai (Liu et al., 2005; Ng et al., 2012), are
reproduced in Fig. 5. The undrained shear strength data from
the Xingye Bank site investigation (Fig. 4) are also included
in this plot.

exhibit creep. It also seems plausible that post-cure mechanisms in any reinforced concrete components (e.g. diaphragm
walls or slabs) will contribute to the tendency of the nearby
ground to exhibit time-dependent movements. Soil creep
effects are excluded in the modelling procedures described
later in this paper, although an attempt is made to incorporate post-cure shrinkage of the floor slabs within the
analysis.
Two alternative soil modelling procedures are available
for the analysis of undrained problems in geotechnical
engineering. One approach, which has been previously used
in the analysis of deep excavations (e.g., Ng & Lings, 1995;
Hashash & Whittle, 1996; Zdravkovic et al., 2005; Kung
et al., 2009), is to adopt an effective stress model for the soil
that is coupled with a nearly incompressible model for the
pore fluid. This approach has the disadvantage, from a practical perspective, that measured data on undrained shear
strength cannot be correlated directly with the model parameters; instead, a separate calibration process is required.
The alternative approach, adopted in the current paper, is to
formulate the soil model as a single phase material in terms
of total stresses. In this case, (approximately) zero volumetric
strains are enforced by way of constraints that are implicit
within the constitutive model. This latter approach has the
considerable advantage, from a practical perspective, that
undrained shear strength is treated as a material parameter;
measured spatial variations of undrained shear strength are
therefore incorporated, in a straightforward way, within the
constitutive model. Moreover, total stress models are in
general more robust computationally than effective stress
models and typically involve significantly less computational
effort. For the detailed analysis presented in this paper,
the robustness of the total stress approach is particularly
advantageous.

Soil layers
0

t: kN/m3
16

18

w n , w l, w p : %
20 20

40

60 05

e
10

Cc
15

05

su: kPa
10

20

40

c: kPa
0

10

Fill
clay
Silty clay

10

Depth below ground level, z: m

Mucky clay

20

Silty clay,
with clay

30
Silty clay,
with clayey
silt
40
wn
wl
50

Sandy silt

wp

Silty clay, with


sandy silt
60 Silty clay, with
silty sand
Note:

t = unit weight, wn = water content, wp = plastic limit, w1 = liquid limit, e = void ratio, Cc = compressive index,
su = field vane shear strength, c = cohesive strength, = internal friction angle

Fig. 4. Geotechnical profile and soil properties from the site investigation (Xu, 2007)

Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

: degrees
0

10

20

30

FINITE-ELEMENT ANALYSIS OF A DEEP EXCAVATION CASE HISTORY


su: kPa
0

50

100

150

250

Dassargues et al. (1991)


Xu (2007)
Ng et al. (2012)
Liu et al. (2005)
Equation (1)

10
Depth below ground level, z: m

200

20

On the basis of the data in Fig. 6, G0 is assumed to increase


linearly with depth. The following correlation is assumed in
the analyses described later in this paper
G0 20 2z

30
40
50
60

su = (20 + 2z) kPa

70

Fig. 5. Undrained shear strength profiles determined from four


separate sites in Shanghai

On the basis that the soil is normally consolidated, or


lightly overconsolidated, it is assumed that the undrained
shear strength increases linearly with depth. The following
variation of shear strength with depth is assumed in the
finite-element analyses described in this paper (where z is
depth in units of metres and su is undrained shear strength in
units of kPa)
su 20 z

Most of the finite-element analyses described later in this


paper are based on the use of a multi-surface kinematic
hardening plasticity model (Houlsby, 1999) to represent the
soil. (Two subsidiary analyses, based on the use of an elastic
perfectly plastic model, have also been conducted for comparison purposes.) To calibrate the kinematic hardening
plasticity model, data are required on the small strain stiffness behaviour, that is, the small strain shear modulus, G0,
and the variation of tangent shear modulus with shear strain.
Appropriate values of G0 can be determined from shearwave velocity tests. Relevant data are given in Cai et al. (2000)
(from the Quyang district of Shanghai), Chen et al. (2011)
(from the site of Shanghai Hongqiao station) and Lou et al.
(2007) (from two further sites in Shanghai). Additional data
on characteristic values of shear wave velocity for depths of
up to 100 m in Shanghai are reported by Gao & Sun (2005).
Data from these sources are plotted in Fig. 6.

Gs
1

G0 1 =05

100

150

Depth below ground level, z: m

10

200

05

30
40

350

250

400

Silty clay, Lu et al. (2005)


Sandy silt, Lu et al. (2005)
Medium sand, Lu et al. (2005)
Huang et al. (2001)
Wang (2004)
Equation (3)

08
07
06
05
04

60

03

70

02

90

09

50

80

1
Ir

10
250

Cai et al. (2000)


Chen et al. (2011)
Gao & Sun (2005)
Lou et al. (2007)
Lou et al. (2007)
Equation (3)

20

where is the shear strain and 05 is a reference strain at


which Gs/G0 05. Similar expressions have been used by
previous researchers (e.g. Hardin & Drnevich, 1972; Stokoe
et al., 1999; Darendeli, 2001; Santos & Correia, 2001). It is
straightforward to show that, for the particular correlation in
equation (3), the reference shear strain is related to rigidity
index by

Gs /G0

50

where the depth z is in units of metres and G0 is in units of


MPa. Equations (1) and (2) imply a rigidity index of Ir 1000
(where Ir G0/su) that is invariant with depth.
It should be noted that there is a potential difficulty in the
use of the relatively simple model for strength and stiffness
given in equations (1) and (2). As indicated in Fig. 1, the
Xingye Bank site is surrounded by a range of existing buildings; the self-weight of these buildings is likely to have
caused additional consolidation in the soil. This is particularly the case at this site where the soil is normally consolidated or lightly overconsolidated. It would be possible,
in principle, to include the effects of these consolidation
processes (which would tend to increase the strength and
stiffness of the soil beneath the neighbouring buildings) in the
model. However, this has not been attempted in the current
analysis.
Only limited data are available in the literature on the
small strain stiffness properties of Shanghai Clay. Lu et al.
(2005) report the results of resonant column tests and cyclic
triaxial tests on three different types of remoulded soil (sandy
silt, silty clay and medium sand). Huang et al. (2001) give
various data determined from triaxial and resonant column
tests. Wang (2004) presents data from bender element tests.
These data, all in terms of secant shear modulus, Gs, normalised by G0, are reproduced in Fig. 7. The data present a
consistent pattern, with the exception of the data from Wang
(2004), which falls below the general trend.
The data in Fig. 7 may be represented, reasonably well, by
the equation

G0: MPa
0

01
G0 = (20 + 2z) MPa = 1000su

100

Fig. 6. Profile of G0 determined from shear-wave velocity tests in


Shanghai

Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

0
106

05
105

104

103

102

101

Shear strain,

Fig. 7. Variation of normalised secant shear modulus Gs/G0 with


shear strain

DONG, BURD AND HOULSBY

Equation (3), with Ir 1000, is plotted in Fig. 7. It is seen


to provide a good fit to the data (with the exception of Wang
(2004)).
The data in Fig. 7 relate to the secant modulus, Gs.
However, to calibrate the multi-surface kinematic hardening
plasticity model adopted for the current analysis, data are
required on the tangent shear modulus, Gt. It is straightforward to show that equation (3) implies a variation of tangent
shear modulus with shear strain of the form

400

10 0 m

Gt
1

G0 1 Ir 2

400

The expression in equation (5) is used to determine the


constitutive parameters for use in the multi-surface kinematic
hardening plasticity model, as described in a later section.

Roller
boundary
y

Roller
boundary

Fixed
boundary

Fig. 9. Geometry and mesh of the whole model

DEVELOPMENT OF THE FINITE-ELEMENT


MODEL
Model description
A finite-element model, developed in Abaqus V611, was
developed to represent the problem at a level of detail that
was judged to be appropriate for the structure and the
various construction processes that were employed in the
project. The geometry and mesh of the model are shown in
Fig. 9. Roller boundary conditions are assigned to the four
vertical sides of the mesh and the bottom is fixed. Initially
a central analysis was established which incorporated best
estimates of the various parameters and procedures that are
needed for the analysis. Subsidiary analyses were then performed to investigate the sensitivity of the results to some of
the assumptions inherent in the central analysis.
The soil is modelled with linear displacement, eightnoded, hexahedral elements with reduced integration

thi

ck

Wall 4

L01

Wall 5

4m

Wall 7

Line 2

Wall 3

P8

Fig. 8. Key field instrumentation locations

Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

312

8m

10 m thick

y z x

Fig. 10. Diaphragm wall geometry and mesh

292 m

Wall 9

l2

AA12
Line 1

Xingye Bank
deep excavation

Wall 6

83

5m
0

AA9

P9

Wall 8

7
87

L06

al
W

Wall 1

(C3D8R). Linear elements were used in this case to facilitate


the development of a model in which the geometry of the
problem could be represented to a high level of detail,
without exceeding the capability of the available computing
resources. It should be noted, however, that linear elements
typically exhibit an over-stiff response when used to conduct
failure analyses in geotechnical engineering (although this
tendency is reduced by the use of reduced integration). The
analyses presented here, however, focus primarily on deformations, for which locking behaviour of the linear elements
should not be significant. In a practical design situation,
consideration should be given to conducting analyses using
higher order elements.
In the central analysis, the diaphragm wall is modelled
using a mesh of C3D8R elements (Fig. 10) with three
elements through the thickness of the wall. In a subsidiary
analysis, an alternative approach, based on the same mesh
topology but employing shell elements, is employed.
The principal structural elements, shown in Fig. 11,
include vertical piles and columns, horizontal beams and
floor slabs. The piles and beams are modelled with linear
beam elements (B31). The floor slabs are modelled with
four-noded quadrilateral shell elements with reduced integration (S4R).
The finite-element mesh used for the central analysis has a
total of 102 036 elements and 116 756 nodes. All analyses
were conducted assuming undrained conditions using a total
stress analysis. The steps used in the analysis follow closely
the construction sequence specified in Table 1. However, the

252 m

Field data
A comprehensive field measurement programme,
described in Xu (2007), was carried out to monitor the
performance of the diaphragm wall and the deformations in
the neighbouring structures, during, and after, the construction process. The numerical calculations presented in the
current paper, are concerned with wall deformations at two
typical points (P8 and P9) (see Fig. 8) where inclinometer
data are available in Xu (2007). Point P9 has been chosen as it
lies at the midpoint of one side of the excavation; point P8 is
located at a re-entrant corner. In addition, computed deformations are presented along two lines on the ground surface
(denoted line 1 and line 2 in Fig. 8) along which settlements,
obtained using optical levels, are reported by Xu (2007).

FINITE-ELEMENT ANALYSIS OF A DEEP EXCAVATION CASE HISTORY

Kinematic hardening model

g1

08
07
06

g2
g3

05
04

Area = 10
in linear plot

g4
c4c3

03
02
01
0
103

102

101

g5
c5c4

where is the stress tensor, J2 is the second invariant of the


deviatoric stresses and C is a parameter that defines the size
of the surface and is interpreted as the undrained shear
strength. In addition, the model includes a set of n inner,
kinematic hardening, yield surfaces with the same shape as
the outer von Mises surface.
The model is specified by the small strain shear modulus, G0,
the bulk modulus K and a set of non-dimensional parameters ci
and gi (i 1, n) that are used to specify the size and work
hardening characteristics of each of the inner surfaces. The size
of each inner surface is ciC and the tangent shear modulus
when the ith surface is active is giG0. A total of nine inner yield

Equation (5)

g0 = 10

c3c2

09

c2c1

f 6J2  8C 0
2

10

c1

Material models and input parameters


The Shanghai Clay is represented by a multi-surface
kinematic hardening soil model (Houlsby, 1999) which has
been developed to represent the small strain non-linear
behaviour of undrained soils. This multi-surface model,
formulated within the framework of work-hardening plasticity theory, is able to represent the non-linear behaviour
of soil at small strains. The model is described in detail by
Houlsby (1999); the use of the model to conduct finiteelement analyses of the deformations around shallow tunnels
is described in Burd et al. (2000). For the current analysis, this
model has been implemented in Abaqus by way of a UMAT
(an interface in Abaqus for users to write user-defined
material constitutive models) subroutine (Dong, 2014).
The soil model consists of a fixed outer von Mises surface
defined by

surfaces are used in the analyses described in this paper, a


balance between accuracy and computational efficiency.
The parameters for the model are selected to provide a fit
with the stiffness degradation curve given in equation (5).
These parameters are listed in Table 2. Details of the procedure used to determine the numerical values of the model
parameters are given in Dong (2014). A comparison between
the step-wise stiffness degradation curve computed using the
multi-surface kinematic hardening plasticity model, using
the data in Table 2, and equation (5) is shown in Fig. 12.
To illustrate the performance of the kinematic hardening
model, an analysis has been conducted of the shearing
phase of a conventional triaxial compression test. The results
of this analysis are plotted in Fig. 13 in terms of normalised

Gt/G0

diaphragm wall installation is modelled as wished-in-place


for simplicity, and the dewatering process is not modelled (on
the basis that a total stress analysis is being conducted).

100

101

102

Ir

Fig. 12. Plot of Gt/G0 against Ir for equation (5) and also for the
kinematic hardening model based on parameters in Table 2

Superstructures

20
Kinematic hardening model

18
Normalised deviator stress, q

Floor slabs
and beams

Equation (3)

16
14
12
10
08
06
04
02
0

05

10

15

20

25

30

35

40

45

50

Normalised triaxial shear strain, s

Fig. 13. Plot of normalised deviator stress q against normalised


triaxial shear strain s . Comparison between the results of triaxial
compression tests determined on the basis of equation (3) and the
performance of the kinematic hardening model (including an unload
reload cycle) based on the parameters used in Table 2. The filled
circles indicate intersections with one of the yield surfaces

z
y

x
Piles and columns, 60 m deep

Fig. 11. Supporting system and superstructures


Table 2. Parameters for the nested kinematic hardening model for n = 9
i

gi
ci

10

09
0005

07
00952

05
02264

03
03537

015
05358

0075
06769

003
07678

00075
08708

000058
0942

Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

DONG, BURD AND HOULSBY

deviator stress, q q=su (where q is deviator stress) and


normalised triaxial shear strain s s Ir (where s is the
increment of triaxial shear strain). In this analysis, the model
is initially loaded to 80% of the failure load (i.e. q 16). The
computed performance of the kinematic hardening model
conforms closely (as expected) to the response computed
using equation (3). At q 16 the model is unloaded and
then re-loaded. The model exhibits a relatively stiff unloading
response and shows typical hysteretic behaviour on reloading.
This is a feature of the model; it is able to capture realistically
not only the change of stiffness on unloading, but also
realistic behaviour on reloading. More generally it captures
the effects of immediate past stress history on the stiffness of
the soil.
The diaphragm wall is modelled using a cross anisotropic elastic model. This provides a means of modelling the
structural influence of the construction joints in the wall
(Zdravkovic et al., 2005). The elastic properties of the

Wall: shell elements


1, continuous, anisotropic
2, discontinuous at corners, isotropic
3, continuous but release rotational
DOFs at corner, isotropic

concrete used to construct the wall are assumed to be Ev


30 GPa, where Ev is the vertical Youngs modulus and
Eh Ev, where Eh is the horizontal Youngs modulus and
specifies the degree of anisotropy. Values of Poisson ratio
are vhv vvh 0.
The horizontal beams and floor slabs are modelled as
isotropic linear elastic materials with Youngs modulus E
30 GPa and Poisson ratio v 02. The floor slabs and beams
act as props for the diaphragm wall. During construction, the
floor slabs are cast against the diaphragm wall. Once a slab
has been cast, however, various complex mechanisms come
into play as the concrete cures (e.g. Kim & Ahn, 2009).
Initially, the concrete will heat up and expand as a consequence of the exothermic curing processes. The slab will
then shrink as it cools. Various other time-dependent shrinkage processes will also occur. This rather complex behaviour
during the curing process has the unfortunate effect that the
precise level of support that the slab provides to the retaining

Conventional soil models


1. Linear elastic, constant properties
2. Tresca, constant properties
3. Tresca, variable properties

Wall: anisotropic, solid


elements, varies
Influence of anisotropic
wall properties

Influence of wall models

Influence of soil models

Central analysis
Soil: multiple-yield surface model
G0 and su increase with depth
= 185 kN/m3, K 0t = 088
Wall: solid element, anisotropic properties,
best = 01
Slab, beam: linear elastic, shrinkage,
= 1 105/K, Tbest = 35 K

Influence of initiaI stress state


K 0t = 077, 10

Influence of concrete shrinkage


T = 30 K, 40 K

Fig. 14. Analysis strategy (DOFs: degrees of freedom)

Table 3. Specification of the subsidiary analyses


Parameter or procedure being explored

Nomenclature

Analysis description

Constitutive model

Central analysis

Multi-surface kinematic hardening plasticity model


su (20 2z) kPa, G0 1000su
Tresca soil model
Uniform strength and stiffness
su 50 kPa, G 125 MPa, v 049
Tresca soil model
Strength and stiffness vary with depth su (20 2z) kPa,
G 250su, v 049
T 35 K
T 30 K
T 40 K
01
105
10
Eight-noded continuum elements, 01
Four-noded shell elements, 01
Kt0 088
Kt0 077
Kt0 10

Tresca 1
Tresca 2
Thermal modelling of slab and beams
Degree of anisotropy of diaphragm wall model
Element type to model diaphragm wall
Horizontal stress distribution

Central analysis
T1
T2
Central analysis
A1
A2
Central analysis
S1
Central analysis
H1
H2

Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

FINITE-ELEMENT ANALYSIS OF A DEEP EXCAVATION CASE HISTORY


wall cannot easily be determined. In addition, the floor slabs
may also shrink or expand due to the variation of ambient
temperature (Whittle et al., 1993; Boone & Crawford, 2000;
Hashash et al., 2003).
Avariety of procedures have been used in previous analyses
to incorporate post-cure concrete shrinkage effects in
finite-element models of propped excavations. One approach
is to use a reduced stiffness for the slab (e.g. Simpson, 1992;
St. John et al., 1993). This approach has the disadvantage,
however, that it does not represent, in any meaningful way,
the detailed physics of the concrete curing process being
modelled. An alternative approach is to model shrinkage
effects by prescribing thermal strains to the floor slabs. In this
approach, thermal and post-cure shrinkage effects are effectively lumped together and dealt with in the model by specifying an appropriate set of thermal strains using a combination
of coefficient of thermal expansion and temperature
change, T. This approach is adopted in the current model
with 105/K. The required amount of thermal shrinkage
is specified by way of an appropriate value of T for the
horizontal beams and slabs.

0
2
4
6
8
10
12
14
16
18
20
22
24
26
28
30
32
40 30 20 10

values of (the wall anisotropy factor) and T (the temperature change required to model post-cure shrinkage in the
horizontal beams and slabs) need to be selected. The values
adopted in the current analysis (chosen on a trial-and-error
basis to provide a reasonable comparison with the field data)
are 01 and T 35 K. All other parameters adopted
in the central analysis were based directly on the available
geotechnical and structural data.
Once the central analysis had been completed, separate
subsidiary calculations were conducted to investigate the
influence of certain key aspects of the model. This process
provides an indication of the sensitivity of the analysis to the
calculation parameters and modelling procedures adopted in
the central analysis.

INTEPRETATION OF RESULTS
Xu (2007) provides a substantial database of field data
that may be compared with the results of the finite-element
analysis. In conducting these comparisons, various issues
need to be considered. First, the horizontal wall movements
reported by Xu (2007) were based on inclinometer readings
and reported on the basis that the displacement at the base
of the inclinometer is zero. To compare these data with the
finite-element results, the computational results have been
shifted to match the zero displacement condition that is
assumed at the base of the inclinometers. In addition, the
data in Xu (2007) indicate that measureable ground

P9
Field data
Central analysis
Tresca 1
Tresca 2

Wall depth: m

Wall depth: m

ANALYSIS STRATEGIES AND CALCULATIONS


A parametric study has been conducted according to the
strategy shown in Fig. 14. The individual calculations that
have been conducted are specified in Table 3. Initially a
central analysis is conducted. For this analysis, appropriate

10

20

30

40

50

60

70

80

90 100

0
2
4
6
8
10
12
14
16
18
20
22
24
26
28
30
32
40 30 20 10

P8
Field data
Central analysis
Tresca 1
Tresca 2

20

30

40

50

60

70

80

90 100

100

60

90

40

Field data
Central analysis
Tresca 1

30

Tresca 2

Vertical ground movement: mm

Line 1

50
Vertical ground movement: mm

10

Wall deflection: mm
(b)

Wall deflection: mm
(a)

20
10
0
10

Line 2
Field data
Central analysis

80
70

Tresca 1
Tresca 2

60
50
40
30
20
10
0
10

20
30

20
0

10

15

20

25

30

35

Horizontal distance from wall: m


(c)

40

30

10

20

30

40

50

60

70

Horizontal distance: m
(d)

Fig. 15. Wall deflections and ground settlements (effects of soil models): wall deflection at (a) P9 and (b) P8; vertical ground movement along
(c) line 1 and (d) line 2

Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

DONG, BURD AND HOULSBY

10

settlements (typically of the order of 5 mm) were induced


during the wall installation and the dewatering processes.
Since the current model does not include these effects,
deformations associated with these construction processes
are excluded from the field data in the comparisons described
below. It should also be noted that diaphragm wall installation is likely to modify the local horizontal stresses in the
ground; the effect that these adjustments in horizontal
stresses might have on the subsequent incremental wall and
ground movements is not considered in the analysis.
The parametric study generated a very substantial amount
of data and only selected results are summarised in this
paper.

0
2
4
6
8
10
12
14
16
18
20
22
24
26
28
30
32

Concrete shrinkage
Two subsidiary analyses with different values of imposed
temperature change T, as indicated in Table 3, have been
used to investigate the influence of induced shrinkage in the
concrete floor slabs and beams on the computed behaviour.

Wall depth: m

Wall depth: m

Influence of soil models


A set of analyses, specified in Table 3, has been conducted
to explore the influence of the choice of soil model on the
computed behaviour. The central analysis uses the multisurface kinematic hardening plasticity model, described
earlier, with a linear variation of strength with depth (equation (1)) and a constant value of rigidity index, Ir 1000. Two
additional calculations have been conducted using the
elasticperfectly cohesive model, based on the Tresca yield
criterion, that is available in the Abaqus material library.
Careful consideration is given to the choice of soil parameters to ensure that the analyses are broadly comparable.
For the Tresca 1 model, the soil properties do not vary with

depth. The strength su is determined from equation (1) at a


depth of 15 m, roughly half of the wall depth, and the
stiffness G is taken to be equal to the tangent stiffness at 50%
of the shear strength for soil at a depth of 15 m. From
equations (1) and (2) this gives G 025G0 125 MPa. For
the Tresca 2 soil model, the shear strength is assumed to vary
linearly with depth, according to equation (1). The shear
modulus varies with depth as G 250su.
Figure 15 shows the calculated wall deflections (at locations
P9 and P8) and the vertical ground movements (along line 1
and line 2). The central analysis appears to capture the wall
deflection and ground movement reasonably well. The Tresca 2
model, captures the overall pattern of wall deflection reasonably well, although it fails to reproduce the pattern and magnitude of the vertical ground movement. The Tresca 1 model,
however, provides a poor comparison with the field data. These
comparisons suggest that, as expected, the kinematic hardening plasticity model is to be preferred over the simpler Tresca
model. Furthermore, comparing Tresca 1 and Tresca 2, it is
clear that modelling the soil as having constant strength and
stiffness with depth is highly unsatisfactory.

P9
Field data
Central analysis
T1
T2

10

15

20

25

30

35

40

45

0
2
4
6
8
10
12
14
16
18
20
22
24
26
28
30
32

50

P8
Field data
Central analysis
T1
T2

10

15

20

25

30

35

40

45

50

Wall deflection: mm
(b)

Wall deflection: mm
(a)
5

0
Line 2

Vertical ground movement: mm

Vertical ground movement: mm

0
5
10
15
Line 1

20

Field data
Central analysis
T1
T2

25
30

10

15

20

25

30

35

Horizontal distance from wall: m


(c)

40

Field data
Central analysis

T1
T2

10

15

20

25

10

20

30

40

50

60

70

Horizontal distance: m
(d)

Fig. 16. Wall deflections and ground movements (thermal effects): wall deflection at (a) P9 and (b) P8; vertical ground movement along (c) line 1
and (d) line 2

Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

FINITE-ELEMENT ANALYSIS OF A DEEP EXCAVATION CASE HISTORY


The results in Fig. 16 show that the calculated wall
deflections and ground movements are sensitive to these
induced shrinkage effects. Increasing the magnitude of the imposed temperature change from T 30 K to T 40 K
has the effect of increasing the computed wall displacements,
as would be expected. The use of a thermal shrinkage model
to represent the various post-cure shrinkage effects that
develop in the floor slabs appears to provide a practical and
plausible approach. However, the choice of the appropriate
shrinkage parameter, T, presents a practical difficulty.
In the current analysis, was set to an assumed value
for the coefficient of thermal expansion of concrete and T
was chosen by comparison between the results of the
finite-element analysis and the available field data on a
trial-and-error basis. For routine design situations, in which
previous field data are unavailable, an alternative approach
will need to be devised to determine an appropriate value of
T for use in the model.

0
2
4
6
8
10
12
14
16
18
20
22
24
26
28
30
32

As shown in Fig. 17, when the isotropic model is


used for the wall ( 10, A2 analysis) the wall deflection
pattern in the finite-element results differs significantly
from the field data at the wall corner, P8, whereas the
moderately anisotropic wall model ( 01, central analysis)
provides a good fit to the field data at this location. As the
anisotropy factor decreases ( 105, A1 analysis), the
magnitude of the wall deflection at P8 increases significantly,
although the bulging pattern is maintained. The deformations at point P9, in the centre of a stretch of wall, are
seen to be less sensitive to the value of than is apparent at
the corner (P8).
The ground settlements along line 1 and line 2, are also
sensitive to the value of . The isotropic wall model ( 10,
A2 analysis) generally underestimates the ground settlement
along line 1 (which is close to the wall corner) while the two
analyses with an anisotropic model for the wall provide a
reasonably good fit to the data. For the ground settlement
along line 2, the settlement close to the wall corner is more
sensitive to the value of than that near the wall centre. This
behaviour is presumably associated with the influence of the
corners in the diaphragm wall.
These observations indicate that the joints between
the wall panels have an important influence on the wall
performance. Careful choice of the anisotropy factor, , is
needed to obtain a satisfactory model. The value 105,
recommended for a contiguous secant piled wall (Zdravkovic
et al., 2005), appears to be too small for the diaphragm wall

Wall depth: m

Wall depth: m

Influence of joints in the diaphragm wall


The influence of construction joints in the diaphragm wall
are included in the analysis by using an anisotropic model for
the wall. The value of the anisotropy factor, , adopted in the
central analysis was determined on a trial-and-error basis.
Subsidiary calculations were conducted to investigate the
sensitivity of the analyses to the value of .

P9
Field data
Central analysis
A1
A2

10

15

20

25

30

35

40

45

11

0
2
4
6
8
10
12
14
16
18
20
22
24
26
28
30
32

50

P8
Field data
Central analysis
A1
A2

10

15

Wall deflection: mm
(a)

20

25

30

35

40

45

50

Wall deflection: mm
(b)

0
Line 2

Vertical ground movement: mm

Vertical ground movement: mm

0
5
10
15
Line 1

20

Field data
Central analysis
A1
A2

25
30

10

15

20

25

30

35

Horizontal distance from wall: m


(c)

40

Field data
Central analysis

A1
A2

10

15

20

25

10

20

30

40

50

60

70

Horizontal distance: m
(d)

Fig. 17. Wall deflections and ground movements (effects of joints): wall deflection at (a) P9 and (b) P8; vertical ground movement along (c) line 1
and (d) line 2

Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

DONG, BURD AND HOULSBY

12

in this case, whereas 01 seems to represent the behaviour


reasonably well.

0
2
4
6
8
10
12
14
16
18
20
22
24
26
28
30
32

Influence of initial horizontal stress distribution


The potential influence of the magnitude of the initial
horizontal stresses in the ground is investigated by conducting subsidiary analyses with different values of the coefficient
of earth pressure at rest, Kt0, based on total stresses. The
central analysis was based on Kt0 088; this value was
determined as follows. For an effective angle of friction of 15
(estimated from the original site investigation data, Fig. 4)
the value of K0 (in terms of effective stresses) determined
from the Jaky formula (K0 1sin) is 0741. This result,
combined with some simple assumptions on the unit weight
of the soil and the pore pressure variation with depth, gives
an estimate of 088 for Kt0. To investigate the sensitivity of the
analysis to variations in the initial horizontal stresses,
subsidiary calculations have been conducted with Kt0 077
and 10 (corresponding to K0 05 and 10). The results are
described below.
The computed wall deflections at P9 and P8 are shown in
Fig. 19. It is clear that changes in Kt0 have an insignificant
influence on the pattern of wall movements. However, the
data indicate that the vertical ground movements along lines
1 and 2 are both sensitive to the value of Kt0. Increasing Kt0

Wall depth: m

Wall depth: m

Choice of shell or solid elements for the wall


The central analysis has been repeated using shell elements
to model the diaphragm wall (rather than solid elements).
The shell element wall has the same cross anisotropic properties as the previous central analysis. Results from the
analyses described in Table 3 are shown in Fig. 18.
As shown in Fig. 18, the shell element wall produces wall
deflection patterns that are similar to those computed using
solid elements, but the displacements are typically about
30% greater in magnitude. Figure 18 also indicates that the
shell element wall results in approximately 30% larger ground
settlement along line 1 and line 2 compared to the solid
element wall in the central analysis.
This finding that computed wall displacements tend to
be greater in magnitude when shell elements are used in the
model, rather than continuum elements, is consistent with
the results of Zdravkovic et al. (2005). Note that this is in
spite of the fact that the bending stiffness of the wall was
matched in the two analyses. Vertical acting shear stresses
develop on the soil/structure interface behind the retaining
wall. When these shear stresses act downwards (as is typically
the case) then additional bending moments are set up the
retaining wall that tend to reduce the magnitude of the wall
deflection. Since the geometric thickness of the shell element

is zero, these additional bending moments are not incorporated in the analysis when shell elements are used. As a consequence, the wall deforms in a more flexible manner when
shell elements are used in the analysis.

P9
Field data
Central analysis
S1

10

15

20

25

30

35

40

45

0
2
4
6
8
10
12
14
16
18
20
22
24
26
28
30
32

50

P8
Field data
Central analysis
S1

10

15

Wall deflection: mm
(a)

20

25

30

35

40

45

50

Wall deflection: mm
(b)

0
Line 2

Vertical ground movement: mm

Vertical ground movement: mm

0
5
10
15

Line 1
Field data

20

Central analysis
S1

Field data

Central analysis
S1

10

15

20

25
30

10

15

20

25

30

35

Horizontal distance from wall: m


(c)

40

25

10

20

30

40

50

60

70

Horizontal distance: m
(d)

Fig. 18. Wall deflections and ground movements (effects of shell elements): wall deflection at (a) P9 and (b) P8; vertical ground movement along
(c) line 1 and (d) line 2

Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

FINITE-ELEMENT ANALYSIS OF A DEEP EXCAVATION CASE HISTORY


tends to increase the magnitude of the computed vertical
ground movements.

CONCLUSIONS
The case study described in this paper suggests that 3D
finite-element analysis is capable of providing realistic data
on the performance of a complex deep excavation. Careful
consideration needs to be given, however, to various aspects
of the model, to ensure that satisfactory results are obtained.
The following conclusions can be drawn from the analyses
presented in this paper.

(b) Post-cure shrinkage effects in the concrete floor slabs


have a significant influence on the structural interaction
between the slab and the retaining walls. It appears
to be possible to develop a satisfactory model for these
shrinkage processes based on a relatively simple thermal
strain approach. Further work is needed to develop
appropriate procedures to determine appropriate thermal parameters for use in routine design.
(c) When shell elements are used to model the retaining
wall, fewer nodes are employed than is the case for a
model based on continuum elements. In spite of their
relative simplicity, however, the use of shell elements in
this application is not recommended. When the retaining wall is modelled with shell elements, the wall deflection and ground settlement are overestimated (by 30%
in this case study) compared with the model with solid
elements for the wall. This difference is associated with
the beneficial effect of downward-acting shear stresses
acting on the back of the retaining all. The effect of
these shear stresses is not modelled correctly when shell
elements are used.
(d) The performance of the retaining wall is influenced by
the presence of construction joints, particularly near the
corners. The current analyses suggest that satisfactory
results can be achieved using anisotropic elasticity to
represent the wall, provided that an appropriate value of
the anisotropy factor, , is adopted. For the diaphragm
wall investigated in the current study a value of 01

P9

0
2
4
6
8
10
12
14
16
18
20
22
24
26
28
30
32

Field data
Central
H1
H2

Wall depth: m

Wall depth: m

(a) An elasticperfectly plastic model was unable to


reproduce the observed patterns of displacement, even
when the strength and stiffness parameters were allowed
to vary with depth. In contrast, the multi-surface
kinematic hardening plasticity model adopted to represent the soil in the central analysis gave results that
conformed closely to the field data. These observations
are consistent with much of the previous work in this
area and confirm the fundamental importance of
adopting a model that is capable of representing small
strain non-linear soil behaviour in order to obtain
realistic results for soil/structure interaction problems of
this sort.

10

15

20

25

30

35

40

45

0
2
4
6
8
10
12
14
16
18
20
22
24
26
28
30
32

50

P8
Field data
Central
H1
H2

10

15

20

25

35

40

45

50

0
Line 2

Vertical ground movement: mm

0
Vertical ground movement: mm

30

Wall deflection: mm
(b)

Wall deflection: mm
(a)

5
10
15
Line 1
Field data
Central analysis
H1
H2

20
25
30

13

10

15

20

25

30

35

Horizontal distance from wall: m


(c)

40

Field data
Central analysis
H1
H2

10

15

20

25

10

20

30

40

50

60

70

Horizontal distance: m
(d)

Fig. 19. Wall deflections and ground movements (Kt0 effects): wall deflection at (a) P9 and (b) P8; vertical ground movement along (c) line 1 and
(d) line 2

Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

DONG, BURD AND HOULSBY

14

appears to be appropriate. The anisotropic properties


that are appropriate for other forms of construction
need to be verified by way of analysis of case histories.
(e) The magnitude of the horizontal stresses in the ground
influences the excavation behaviour. The computed
results indicate that the wall deflection pattern is
insensitive to changes in the value of Kt0. However, the
value of Kt0 does have a significant influence on the
computed ground surface settlements.

ACKNOWLEDGEMENTS
The first author was supported by the China Scholarship
Council to study at Oxford University. The field measurements were conducted by Dr Z. H. Xu, who also analysed the
initial data. The calculations were conducted at the Oxford
Supercomputing Centre.
NOTATION
C
ci, gi
Eh
Ev
G
G0
Gs
Gt
Ir
i
J2
K
K0
Kt0
n
q
q
su
z

T
s
s

05
, hv, vh

parameter in soil model to define size of outmost yield


surface
non-dimensional normalised parameters in soil model
to define size and stiffness associated with each inner
yield surface
Youngs modulus in horizontal direction
Youngs modulus in vertical direction
shear modulus
shear modulus at very small strain
secant shear modulus
tangent shear modulus
rigidity index
ith inter yield surface of the soil model
second invariant of deviatoric stresses
bulk modulus in soil model
coefficient of lateral earth pressure at rest in effective
stress expression
coefficient of lateral earth pressure at rest in total stress
expression
number of yield surfaces used in soil model
deviator stress
normalised deviator stress
undrained shear strength of soil
depth below ground
coefficient of thermal expansion
degree of anisotropy of diaphragm wall
temperature change
increment of triaxial shear strain
normalised triaxial shear strain
effective friction angle of soil
engineering shear strain
reference shear strain at Gs/G0 05
Poisson ratio
stress tensor

REFERENCES
Boone, S. J. & Crawford, A. M. (2000). Braced excavations: temperature, elastic modulus, and strut loads. J. Geotech.
Geoenviron. Engng 126, No. 10, 870881.
Burd, H. J., Houlsby, G. T., Augarde, C. E. & Liu, G. (2000).
Modelling tunnelling-induced settlement of masonry buildings.
Proc. Instn Civ. Engrs Geotech. Engng 143, No. 1, 1729.
Cai, H., Zhou, J. & Li, X. (2000). Plastoelastic response of
horizontally layered sites under multi-directional earthquake
shaking. Tongji Daxue Xuebao/J. Tongji University 28, No. 2,
177182.
Chen, Q. S., Gao, G. Y. & Yang, J. (2011). Dynamic response of deep
soft soil deposits under multidirectional earthquake loading.
Engng Geol. 121, No. 12, 5565.
Darendeli, M. B. (2001). Development of a new family of normalized
modulus reduction and material damping curves. PhD thesis,
The University of Texas at Austin, Austin, TX, USA.

Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

Dassargues, A., Biver, P. & Monjoie, A. (1991). Geotechnical


properties of the Quaternary sediments in Shanghai. Engng
Geol. 31, No. 1, 7190.
Dong, Y. (2014). Advanced finite element analysis of deep excavation
case histories. DPhil thesis, University of Oxford, Oxford, UK.
Finno, R. J. & Bryson, L. S. (2002). Response of building adjacent
to stiff excavation support system in soft clay. J. Perf. Constr.
Facilities 16, No. 1, 1020.
Finno, R. J. & Nerby, S. M. (1989). Saturated clay response
during braced cut construction. J. Geotech. Engng 115, No. 8,
10651084.
Finno, R. J., Atmatzidis, D. K. & Perkins, S. B. (1989). Observed
performance of a deep excavation in clay. J. Geotech. Engng 115,
No. 8, 10451064.
Gao, F. & Sun, X. (2005). Characteristic analysis of Shear wave
velocity of foundation ground in Shanghai region. Shanghai
Geol. 94, No. 2, 2729.
Gourvenec, S., Powrie, W. & De Moor, E. K. (2002).
Three-dimensional effects in the construction of a long
retaining wall. Proc. Instn Civ. Engrs Geotech. Engng
155, No. 3, 163173, http://dx.doi.org/10.1680/geng.2002.155.
3.163.
Hardin, B. O. & Drnevich, V. P. (1972). Shear modulus and damping
in soils: measurement and parameter effects. ASCE J. Soil
Mech. Found. Div. 98, No. SM6, 603624.
Hashash, Y. M. A. & Whittle, A. J. (1996). Ground movement
prediction for deep excavations in soft clay. J. Geotech.
Geoenviron. Engng 122, No. 6, 474486.
Hashash, Y. M. A., Marulanda, C., Kershaw, K. A., Cording, E. J.,
Druss, D. L., Bobrow, D. J. & Das, P. K. (2003). Temperature
correction and strut loads in Central Artery excavations.
J. Geotech. Geoenviron. Engng 129, No. 6, 495505.
Houlsby, G. T. (1999). A model for the variable stiffness of
undrained clay. Proceedings of the international symposium on
pre-failure deformations of soil, Torino, Italy, pp. 443450.
Huang, Y., Brown, S. F. & McDowell, G. R. (2001). Dynamic
coupled analysis for rutting in flexible pavement foundations
under cyclic loading. Yantu Gongcheng Xuebao/Chin. J.
Geotech. Engng 23, No. 6, 757.
Kim, H. & Ahn, H. (2009). Shrinkage stress analysis of concrete
slabs in a multi-storey building considering internal and
external restraints. Struct. Des. Tall Special Buildings 18,
No. 5, 525537.
Kung, G. T. C., Ou, C. Y. & Juang, C. H. (2009). Modeling
small-strain behavior of Taipei clays for finite element analysis of
braced excavations. Comput. Geotech. 36, No. 12, 304319.
Lee, F. H., Hong, S., Gu, Q. & Zhao, P. (2011). Application of
large three-dimensional finite-element analyses to practical
problems. Int. J. Geomech. 11, No. 6, 529539.
Liu, G. B., Ng, C. W. W. & Wang, Z. W. (2005). Observed
performance of a deep multistrutted excavation in Shanghai
soft clays. J. Geotech. Geoenviron. Engng 131, No. 8, 10041013.
Liu, G. B., Jiang, R. J., Ng, C. W. W. & Hong, Y. (2011).
Deformation characteristics of a 38 M deep excavation in soft
clay. Can. Geotech. J. 48, No. 12, 18171828.
Lou, M., Li, Y. & Li, N. (2007). Analyses of the seismic responses of
soil layers with deep deposits. Frontiers Archit. Civ. Engng in
China 1, No. 2, 188193.
Lu, X., Li, P., Chen, B. & Chen, Y. (2005). Computer simulation
of the dynamic layered soil-pile-structure interaction system.
Can. Geotech. J. 42, No. 3, 742751.
Ng, C. W. W. & Lings, M. L. (1995). Effects of modeling
soil nonlinearity and wall installation on back-analysis of
deep excavation in stiff clay. J. Geotech. Engng 121, No. 10,
687695.
Ng, C. W. W., Hong, Y., Liu, G. B. and Liu, T. (2012). Ground
deformations and soilstructure interaction of a multi-propped
excavation in Shanghai soft clays. Gotechnique 62, No. 10,
907921, http://dx.doi.org/10.1680/geot.10.P.072.
Potts, D. & Zdravkovic, L. (2001). Finite element analyisis in
geotechnical engineering: application, vol. 2. London, UK:
Thomas Telford.
Potts, D. M. & Fourie, A. B. (1984). Behaviour of a propped
retaining wall: results of a numerical experiment. Gotechnique
34, No. 3, 383404, http://dx.doi.org/10.1680/geot.1984.34.3.
383.

FINITE-ELEMENT ANALYSIS OF A DEEP EXCAVATION CASE HISTORY


Santos, J. A. & Correia, A. G. (2001). Reference threshold
shear strain of soil. Its application to obtain an unique
strain-dependent shear modulus curve for soil. Proceedings
of the 15th international conference on soil mechanics and
geotechnical engineering, Istanbul, Turkey, vol. 13,
pp. 267270.
SGIDI (Shanghai Geotechnical Investigation & Design Institute)
(1997). Geotechnical site investigation report of Shanghai Xingye
Bank Building, Report no. 97-G-047. Shanghai, China:
Shanghai Geotechnical Investigation & Design Institute.
Simpson, B. (1992). Retaining structures: displacement and design.
Gotechnique 42, No. 4, 541576, http://dx.doi.org/10.1680/
geot.1992.42.4.541.
Skempton, A. W. & Ward, W. H. (1952). Investigations concerning a
deep cofferdam in the Thames Estuary clay at Shellhaven.
Gotechnique 3, No. 3, 119139, http://dx.doi.org/10.1680/geot.
1952.3.3.119.
St John, H. D., Potts, D. M., Jardine, R. J. & Higgins, K. G. (1993).
Prediction and performance of ground response due to
construction of a deep basement at 60 Victoria Embankment.
Predictive soil mechanics. Proceedings of the Wroth memorial
symposium, Oxford, UK, pp. 581608.
Stokoe, K. H., Darendeli, M. B., Andrus, R. D. & Brown, L. T.
(1999). Dynamic soil properties: laboratory, field and correlation studies. Earthquake geotechnical engineering. Proceedings

Downloaded by [] on [11/12/16]. Copyright ICE Publishing, all rights reserved.

15

of the 2nd international conference on earthquake geotechnical


engineering, Lisbon, Portugal, vol. 3, pp. 811845.
Tan, Y. & Wei, B. (2012). Observed behaviors of a long and
deep excavation constructed by cut-and-cover technique in
Shanghai soft clay. J. Geotech. Geoenviron. Engng 138, No. 1,
6988.
Wang, Z. W. (2004). Research on deformation and earth pressure of
deep metro excavation in soft clay based on time and small strain.
PhD thesis, Tongji University, Shanghai, China.
Wang, Z. W., Ng, C. W. W. & Liu, G. B. (2005). Characteristics
of wall deflections and ground surface settlements in Shanghai.
Can. Geotech. J. 42, No. 5, 12431254.
Whittle, A. J., Hashash, Y. M. A. & Whitman, R. V. (1993). Analysis
of deep excavation in Boston. J. Geotech. Engng, ASCE 119,
No. 1, 6990.
Wood, L. A. & Perrin, A. J. (1984). Observations of a strutted
diaphragm wall in London Clay: a preliminary assessment.
Gotechnique 34, No. 4, 563579, http://dx.doi.org/10.1680/
geot.1984.34.4.563.
Xu, Z. H. (2007). Deformation behaviour of deep excavations
supported by permanent structure in Shanghai soft deposit.
PhD thesis, Shanghai Jiao Tong University, Shanghai, China.
Zdravkovic, L., Potts, D. M. & St. John, H. D. (2005). Modelling
of a 3D excavation in finite element analysis. Gotechnique 55,
No. 7, 497513, http://dx.doi.org/10.1680/geot.2005.55.7.497.

You might also like