You are on page 1of 23

INTERNATIONAL JOURNAL FOR NUMERICAL AND ANALYTICAL METHODS IN GEOMECHANICS

Int. J. Numer. Anal. Meth. Geomech., 2005; 29: 49–71


Published online in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/nag.404

Time-dependent modelling for soils and its


application in tunnelling

A. Purwodihardjo and B. Camboun,y


LTDS, Ecole Centrale de Lyon, UMR CNRS 5513, BP 163, 36 Av Guy de Collongue, 69134 Ecully, France

SUMMARY
Monitoring of the progressive convergence of a tunnel shows that deformations occurring in the soil
surrounding a tunnel exhibit a strong evolution with time. This time-dependent behaviour can be linked to
three essential factors: the distance from the point of interest to the working face over time, the distance of
unsupported tunnel to the working face and the viscous properties of the soil.
The objective of this paper is to propose a constitutive model of the time-dependent behaviour of soil
which has been developed within the framework of elastoplasticity–viscoplasticity and critical state soil
mechanics. The consideration of viscoplastic characteristic sets the current model apart from the CJS
(Cambou, Jafari and Sidoroff) model as the basic elastoplastic model, and introduces an additional viscous
mechanism. The evolution of the viscous yield surface is governed by a particular hardening called ‘viscous
hardening’ with a bounding surface.
The proposed constitutive model has been applied in the analysis of tunnelling. Two kinds of numerical
calculations have been used in the analysis, axisymmetric analysis and plane strain analysis. Monitoring of
the progressive convergence of a tunnel conducted in the railway tunnel of Tartaiguille (France), has been
used to describe the calculation procedure proposed and the capability of the model.
The finite difference software, fast Lagrangian analysis of continua (FLAC), has been used for the
numerical simulation of the problems. The comparison of results shows that the observed deformations
could have been reasonably predicted by using the constitutive model and calculation strategy proposed.
Copyright # 2004 John Wiley & Sons, Ltd.

KEY WORDS: viscoplasticity; time dependent; strain softening; CJS; tunnel; numerical model

1. INTRODUCTION

The behaviour of a tunnel is greatly influenced by the characteristics of the soils and the
tunnelling procedure. They will strongly influence the initial and long-term deformations in the
vicinity of a tunnel and on the ground surface, particularly when the ground traversed by
tunnels has poor geotechnical characteristics: little or no cohesion, medium to high
deformability and viscous characteristics. In this area, more consideration should be taken

n
Correspondence to: Bernard Cambou, LTDS, Ecole Centrale de Lyon, UMR CNRS 5513, BP 163, 36 Avenue Guy de
Collongue, 69134 Ecully, France.
y
E-mail: bernard.cambou@ec-lyon.fr

Received 15 March 2004


Copyright # 2004 John Wiley & Sons, Ltd. Revised 9 September 2004
50 A. PURWODIHARDJO AND B. CAMBOU

because deformation of the soil on the ground surface and in the vicinity of a tunnel show
generally a strong evolution with time. This evolution is essentially related to three influences:

* the distance from the point of interest to the working face over time,
* the distance of unsupported tunnel to the working face,
* the viscous properties of the soil.

A good understanding of the above influences as well as an understanding of the effects of the
tunnel support design and installation is necessary in order to predict the deformations induced
by tunnelling. Therefore, a time-dependent model that considers the above influences has been
developed within the framework of elastoplasticity–viscoplasticity. It is based on the CJS
(Cambou, Jafari and Sidoroff) elastoplastic model but also includes an additional viscous
mechanism.

2. DESCRIPTION OF THE MODEL

The CJS model (Cambou, Jafari and Sidoroff) is a constitutive model with different hierarchical
levels which has been developed over the last 20 years by Cambou et al. [1]. This model is based
on non-linear elasticity and two mechanisms of plasticity. It also takes into account the
dependency on density of geomaterials through the critical state [2]. The rate of the strain tensor
can be decomposed into an elastic part and a plastic part. The plastic deformations are
generated from an isotropic and a deviatoric mechanism. Figure 1 shows the two yielding
surfaces linked to these mechanisms in the CJS model.
The total strain in the model is then decomposed in four parts:

e’ ij ¼ e’eij þ e’ip ’dp


ij þ e ’ vp
ij þ e ij ð1Þ

The first part is the elastic part, the second is linked to an isotropic plastic mechanism, the third
part is linked to a deviatoric plastic mechanism and the last part is concerned with an added
viscous mechanism.

Figure 1. Yield surfaces linked to plastic deviatoric mechanism and plastic isotropic mechanism in CJS.

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
TIME-DEPENDENT MODELLING FOR SOILS 51

2.1. Brief description of the basic elastoplastic model of CJS


As it has been explained previously, this model consists of three major mechanisms, i.e. an
elastic mechanism, an isotropic plastic mechanism and a deviatoric plastic mechanism.

2.1.1. Elastic mechanism. The elastic law is given by the following incremental non-linear
equation:
s’ij I’1
e’eij ¼ þ ð2Þ
2G 9K
with
I1 ¼ skk dij

I1
sij ¼ sij 
3
I1 and sij are the first invariant and the deviatoric part of the stress tensor, respectively. dij is the
Kroenecker’s delta symbol. K and G are the tangent bulk and shear modulus, respectively,
which depend on the stress state through a power law:
   
e I1 n I1 n
K ¼ K0 and G ¼ G0 ð3Þ
3Pa 3Pa
K0e ; G0 and n are model parameters while Pa is the atmospheric pressure which equals 100 kPa:

2.1.2. Isotropic plastic mechanism. The yield surface associated with this mechanism is a plane
perpendicular to the hydrostatic axis. The yield surface is given by
I1
f i ðIi ; QÞ ¼  ðQ þ Tr Þ ¼ 0 ð4Þ
3
The yield surface’s evolution is defined by an isotropic hardening mechanism depending on a
scalar variable Q and Tr is a parameter of the model to take into account the cohesion. The
hardening rule has the form
 n
Q
Q’ ¼ K p r’ ¼ K0p r’ ð5Þ
Pa
The isotropic flow rule is described as
9
@f i i>
e’ ip
v ¼ 3li
¼l >
>
>
@I1 =
e’ip
v ¼ r’ ð6Þ
>
>i
@f >
r’ ¼ li ¼ li >
;
@Q
K0p is the tangent plastic bulk modulus and n is a parameter of the model which can be
determined by experiment. li is a parameter of magnitude for the isotropic plastic mechanism
and r is the hardening parameter for this mechanism.

2.1.3. Deviatoric plastic mechanism. For the sake of simplicity no kinematic hardening but only
isotropic hardening was taken into account in the model used in this paper. The yield surface

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
52 A. PURWODIHARDJO AND B. CAMBOU

can be defined by the following equation:


f d ðsij ; RÞ ¼ sII hðyÞ  RðI1 þ 3Tr Þ ¼ 0 ð7Þ

with
 pffiffiffiffiffi sIII 1=6
hðyÞ ¼ 1  54 g
ðsII Þ3
pffiffiffiffiffiffiffiffiffi
sII ¼ sij sij

sIII ¼ detðsij Þ

where g is a parameter of the model and Tr is a parameter of the model to take the cohesion
into account. sII and sIII are the second and the third invariants of the deviatoric stress,
respectively.
The evolution of the yield surface is characterized by the evolution of R with the internal
variable p: The relationship between R and p is written as
ARm p
R¼ ð8Þ
Ap þ Rm
where Rm is a parameter that corresponds to the mean radius of the rupture surface and A is a
parameter of the model. The evolution of (the hardening parameter) p is defined by
   
@f d I1 1:5 I1 1:5
p’ ¼ ld ¼ ld I 1 ð9Þ
@R 3Pa 3Pa
The deviatoric flow rule is given by
 d  d  
@gd d @f @f
e’ dp
ij ¼l d
¼l  nkl nij ð10Þ
@sij @sij @skl
where ld is a magnitude parameter of the incremental plastic strain for the deviatoric
plastic mechanism. The deviatoric potential function ðgd Þ used in Equation (10) corresponds
to a non-associated plastic mechanism. Tensor nij is a symmetrical tensor so that trðn2 Þ ¼ 1
and it is a tangential tensor to the surface corresponding to the potential function. It is
defined by
 
b0 ðsij =sII Þ  dij sII
nij ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffi with b0 ¼ sgnðsij e’dp
ij Þb  1 ð11Þ
02 scII
b þ3
where scII represents the second invariant of the deviatoric stress in the characteristic state and b
is a parameter of the model. The characteristic surface is defined by
f c ¼ scII hðyÞ  Rc ðI1 þ 3Tr Þ ¼ 0 ð12Þ
where Rc represents the mean radius of the characteristic surface. The characteristic
surface corresponds to stress states with no volume change. The rupture surface is the
locus of stress states corresponding to the peak of the stress–strain curve in standard
triaxial tests (standard strain rate). Figure 2 shows the deviatoric mechanism in the
CJS model.

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
TIME-DEPENDENT MODELLING FOR SOILS 53

Figure 2. Different surfaces in the deviatoric mechanism of the CJS model.

2.1.4. Critical state. Two important phenomena can be cited from the drained triaxial tests:
* an increasing of the peak resistance with the initial density of the material,
* the material tends to the state called the critical state characterized by a null volume
variation and a ratio q=p constant independent of the initial density.
To take into account these phenomena in this model, the radius of the rupture surface varies
as a function of the mean effective stress and the density of the material. To simplify the
problem, we take the critical state similar to the characteristic state. Hence, the evolution of the
rupture surface is defined by
 
pc
Rm ¼ Rc þ m ln 0 ð13Þ
p
where m is a parameter of the model, p0 is a mean stress and pc is a critical pressure which is
defined by
pc ¼ pc0 expðcev Þ ð14Þ
with
e  e0
Dev ¼
1 þ e0
c is a parameter of the model, pc0 is a critical pressure corresponding to the initial density, ev is
the accumulated volume strain, e and e0 are the actual and the initial void ratio, respectively.

2.1.5. Strain softening model. In the original CJS model, the definition of the critical state
allows the strain softening of soils which is linked to the dilatancy of soils to be taken into
account. But in the other case, such as for the overconsolidated clay, the strain softening of the
material is generally linked to the cracking which does not develop a significant evolution of the
density. In this case, it is necessary to define another strain softening mechanism which is not
directly linked to the volume strain. So in this paper, the proposed model takes into account the
strain softening of soils which depends on the accumulated deviatoric strain. This model is made
up of three portions, an elastoplastic portion up to the peak strength, a softening portion in
which the strength (Rc and Tr Þ reduces from the peak to residual, and finally, a constant residual
strength portion. Many authors [3–7] have used this type of model. In this model after the peak
strength, the strength parameters, i.e. Rc and Tr will be defined as functions of the second
invariant of the accumulated deviatoric strain ðeII Þ: The following equations are used to define

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
54 A. PURWODIHARDJO AND B. CAMBOU

the softening of the strength parameters:


8
< Rc eff ¼ Rc peak
if eII 4e0II
:
Tr eff ¼ Tr peak
8 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
>
>
> ðe f  eII Þ2
> Rc eff ¼ Rc peak  ðRc peak  Rc res Þ 1  IIf
>
>
< ðeII  e0II Þ2
if e0II 5eII 4eIIf sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð15Þ
>
>
>
> ðeIIf  eII Þ2
>
> T ¼ T  ðT  T Þ 1 
: r eff r peak r peak r res
ðeIIf  e0II Þ2
8
< Rc eff ¼ Rc res
if eII 4e0II
:
Tr eff ¼ Tr res
with
pffiffiffiffiffiffiffiffiffi
eII ¼ eij eij

ekk dij
eij ¼ eij 
3
where the subscripts eff, res and peak refer to an effective value, peak value and residual value,
respectively. In this model, it was assumed that the characteristic state value is the same as the
critical state value. Figure 3 shows the second kind of strain softening behaviour in the CJS
model.

2.2. Viscous hardening with a bounding surface


The viscous effect of the soil is taken into account through a creep surface which is bounded by
the current (artificial) state of stress surface. It means that the creep surface can evolve but the
evolution is limited by the state of stress surface. The current state of stress surface is a surface
homothetic to the yield surface. Meanwhile, the evolution of state of stress surface is limited by
the yield surface (elastoplastic concept). The evolution of the yield surface is limited by the
rupture surface. Figure 4 shows an illustration of the viscous evolution concept with a bounding
surface.

s II

e II
e II0 e IIf
Figure 3. Second kind of strain softening in the CJS model.

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
TIME-DEPENDENT MODELLING FOR SOILS 55

s11 Rupture surface (Rm)

Yield surface (R)

Current stress state surface (Re)

State of stress
Creep surface (Rv)

s22 s33
Figure 4. Illustration of viscous evolution concept with a bounding surface.

The basic formulation for this viscous mechanism is inspired by the overstress model of
Perzyna [8]. To keep within the framework of elastoplasticity is the reason for using this
formulation. The idea is, therefore, starting from the general framework of elastoplasticity and
introducing a modelling of viscous effects.
Many authors [9–13], have employed this formulation, and they have shown that this model is
incapable for introducing the accelerated deformation phenomenon in the case of tertiary creep.
Therefore, in this paper a formulation has been proposed to allow the tertiary creep to be taken
into account in the model.
Thus, three important terms have to be defined in the framework of this model. The first one
is the viscosity of the material, the second one is the function of retardation and the last one is
the direction of the viscoplastic strain. The evolution of the viscous strain is as follows:
1
e’ vd m vd
ij ¼ ðFv Þ Gij ð16Þ
Z
where Z is the viscosity of the material, Fv is the function of retardation and Gvd
ij is the direction
of the viscoplastic strain.
The viscosity of the material in this model is a function of the distance of the state of stress
ðRe Þ from the rupture surface ðRm Þ: This function is defined by
 2 !k
Re
Z ¼ Z0 1  ð17Þ
Rm

where Z0 is a parameter of the model, Re is the mean radius of the current state of stress surface,
Rm is the mean radius of the rupture surface and k is a parameter of the model.
The function of retardation, Fv ; is inspired by the bounding surface theory [14]. This function
is defined from
Re  Rv
Fv ¼ ð18Þ
Rm  Rv

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
56 A. PURWODIHARDJO AND B. CAMBOU

where Rv corresponds to the mean radius of the creep surface. The power m in Equation (16) is
defined as
Re
m ¼ m1 þ m2 ð19Þ
Rm
where m1 and m2 are parameters of the model. The definition of Equation (18) has been selected
for the sake of simplicity in the measurement of the distance in the stress space. The function of
retardation, Fv ; takes an important role because it will drive the evolution of the primary creep,
the secondary creep and the tertiary creep where the secondary creep will be reduced to a
passing point between the primary creep and the tertiary creep. The primary creep, in the creep
test, can be modelled when the creep surface increases and approaches the current state of stress
surface. In this case, the viscous strain rate decreases by the increasing of the creep surface. Two
phenomena can generate the tertiary creep:
(a) the state of stress surface is near to the rupture surface,
(b) the softening of the material resistance (strain-softening behaviour).
In phenomenon type (a) the current state of stress surface is very close to the rupture surface so
the value of Fv is equal to unity (1.0), furthermore, Equation (17) becomes zero and Equation
(16) becomes infinite (rupture). In phenomenon type (b) the rupture surface decreases by
softening of the material resistance and approaches the current state of stress so the value of Fv
is also equal to unity (1.0).
Akai et al. [15], Lade and Liu [16] have shown from laboratory test results that the potential
plastic surfaces for the elasto-viscoplastic and the elastoplastic are homothetic. Based on this
idea, it means that the direction of the plastic strain in the elastoplastic is similar to the
viscoplastic one. Thus, the direction of viscoplastic strain is defined as
 
@fe @fe
Gvd
ij ¼  n kl nij ð20Þ
@sij @skl

where fe is the state of stress surface which is homothetic to the yield surface for the deviatoric
mechanism. It is defined by
f e ¼ sII hðyÞ  Re ðI1 þ 3Tr Þ ¼ 0 ð21Þ

The rupture surface is defined by


f r ¼ sII hðyÞ  Rm ðI1 þ 3Tr Þ ¼ 0 ð22Þ

The creep surface is defined by


f v ¼ sII hðyÞ  Rv ðI1 þ 3Tr Þ ¼ 0 ð23Þ

The evolution of the creep surface is given by


0 1
B 1 C
Rv ¼ Re B
@1   C ð24Þ
Rm vd A
exp Av e
Re II

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
TIME-DEPENDENT MODELLING FOR SOILS 57

Av is a parameter of the model and evd


II is the accumulated deviatoric viscoplastic strain, which is
defined by
qffiffiffiffiffiffiffiffiffiffiffiffi
e’vd
II ¼ e’vd
ij e’ij
vd
ð25Þ

where e’vd
ij is the deviatoric viscoplastic strain rate.

2.3. Parameter identification


The identification of elastoplastic mechanism parameters can be determined by using classical
laboratory tests. The procedure for calibrating the model parameters is briefly defined by Maleki
[2]. Creep tests with different levels of stress have to be achieved, for identifying the viscoplastic
mechanism parameters. This procedure has been explained in detail by Purwodihardjo [17].

3. VALIDATION OF THE MODEL FROM LABORATORY TESTS

The results of laboratory tests performed by Piepi [18] have been used to illustrate the ability of
the model modelling the strain-softening and the time-dependent behaviour of the soils.
Piepi [18] has carried out laboratory tests on clays from Aisne (France). In this validation, a
horizontal layer called Callovo–Oxfordien has been used. This layer with 153 m of thickness is
located at 325–478 m of depth from the ground surface. The average water content of this layer
is 9.1% and the average content of carbonate is 17%. The majority of this layer is composed of
marls (70%). The mean specific gravity of this layer is 2.23, the average porosity is less than 20%
and the average degree of saturation is 96%. Based on the previous physical characteristics,
Rousset [19] has classified this layer as stiff clay.
First, the elastoplastic parameters are identified by using drained triaxial tests performed on
samples taken from 407:6 m depth. The water content is 10.7% and the specific gravity is 2.14.
The elastoplastic parameters identified are listed in Table I.
By using the above parameters, numerical simulations are performed and are compared with
experimental results (Figure 5). It can be seen from this figure that the model can describe quite
satisfactorily the strain softening behaviour of the soils.
After identifying the elastoplastic parameters, the viscoplastic parameters of the model can be
identified. For that, a series of undrained creep triaxial tests have been carried out on samples
taken from 377:19 m depth. The water content of these samples is 9%, the content of carbonate
is 13% and the specific gravity is 2.05. The confinement pressure applied in these tests is 5 MPa:
The loading used on the samples is stepped. Fours levels of loading have been used, i.e. q ¼
9; 10; 11 and 12 MPa: The viscoplastic parameters have been identified by using the first three
levels of loading (9,10 and 11 MPa), while the last level of loading has been used for validating
the model until rupture (tertiary creep or accelerated creep). The viscoplastic parameters
identified can be seen in Table II.
From Figure 6, we can see that the constitutive model proposed can describe very well the
undrained creep process including accelerated creep and creep rupture.
By using the same parameters identified previously, a numerical simulation of a creep triaxial
test from a sample taken from 414:62 m depth is achieved. This sample has been chosen because
the physical parameters are very similar to those of the previous sample (377:19 m depth). In
this validation, the level of loading and confinement pressure used are different from the

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
58 A. PURWODIHARDJO AND B. CAMBOU

Table I. Elastoplastic parameters.


Parameter Value
G0 (MPa) 82.59
K0e (MPa) 79.83
K0p (MPa) 79.83
n 0.6
g 0.22
A ðkPa1 Þ 0.75
Rr up 0.061
Rc 0.055
Rc residual 0.055
b 0.1575
m 0.02
c 90.0
Pc0 (MPa) 17.75
m0 5.0
Tr peak (MPa) 50.05
Tr residual (MPa) 34.33
e0II 0.025
eIIf 0.16

Figure 5. Numerical simulation of drained triaxial tests, w ¼ 10:7%; d ¼ 2:14; depth ¼ 407:6 m (identi-
fication tests; experiments from Piepi [18]): (a) s3 ¼ 5 MPa; and (b) s3 ¼ 10 MPa:

previous sample. The confinement pressure used is 8 MPa and the deviatoric stress applied is
8 MPa: The results of the numerical simulation are presented in Figure 7. From this figure, we
can see that the results of the numerical simulation match the experimental results.

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
TIME-DEPENDENT MODELLING FOR SOILS 59

Table II. Viscoplastic parameters.


Parameter Value
Av 100.0
Z0 ðsÞ 107:89
k 20.0
m1 0.0
m2 6.75

Figure 6. Numerical simulation of undrained stepped loading creep triaxial tests, w ¼ 9%; d ¼ 2:05;
Ca ¼ 13%; depth ¼ 377:19 m (identification and validation tests; experiments from Piepi [18]).

We also validated this model using relaxation tests. Figure 8 compares the undrained stress
relaxation response of the clays from Aisne (France) with a water content 10.26%, a specific
gravity equal to 2.22 and a carbonate content equal to 20.4%. The sample used in this test is
taken from 413:58 m depth. Three levels of strain have been applied to the samples, i.e. 0.25, 0.9,
1.3 and 1.8%. Two types of confinement pressure have been used, i.e. 1 and 10 MPa: It can be
seen that the results of the simulation are qualitatively in agreement with experimental results
where the relaxation gradients are quite similar. But for the second relaxation step, the stress
levels are quite different. It can be noticed that this relaxation test occurs after a change in the
confinement pressure from 1 to 10 MPa: The carbonate content of this sample is higher than the
sample used for the identification of elastoplastic parameters. That is the reason why the
evolution of the plastic modulus with isotropic stress seems, in particular, to be too small in the
simulation with respect to experimental data.

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
60 A. PURWODIHARDJO AND B. CAMBOU

Figure 7. Numerical simulation of undrained creep triaxial tests, w ¼ 11:07%; d ¼ 2:23; CaCO3 ¼ 13:78
%; depth ¼ 414:62 m (validation test; experiments from Piepi [18]).

Figure 8. Numerical simulation of undrained relaxation triaxial tests, w ¼ 10:26%; d ¼ 2:22;


Ca ¼ 20:4%; depth ¼ 413:58 m (validation test; experiments from Piepi [18]).

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
TIME-DEPENDENT MODELLING FOR SOILS 61

4. TIME-DEPENDENT ANALYSIS IN THE TUNNELLING

After the proposed model has been validated on several laboratory tests results including the
results presented in the previous section, this model has been used for the numerical analysis of
tunnelling. To analyse the ability of the model in actual work conditions, data obtained from the
tunnel of Tartaiguille have been considered.

4.1. Presentation of the case study


The tunnel of Tartaiguille is located on the new high-speed train line between Valence and
Mont!elimar (France). This tunnel is double track single tube, allowing a speed of 300 km=h: It
crosses fractured limestones on the north side, stiff marls and sandstones in the south and
stampian clays in the central parts.
From the cross section of the tunnel, it can be seen (Figure 9) that the tunnel support is
shotcrete ðthickness ¼ 300 mmÞ with a steel frame every 1:5 m: Five samples of soil blocks have
been obtained at the section PM 1168 by CETU (Centre d’Etudes des Tunnels) during
construction processes in this tunnel and these blocks have been analysed by Serratrice [20]
(Laboratoire Regional des Ponts et Chausses d’Aix en Provence). From the five samples of soil
blocks, three layers of soil can be concluded at that section. The upper one is the black marl, the
middle one is the calcareous marl and the lower one is the grey marl. The soil characteristics for
the black marl and the grey marl are almost the same, on the other hand the soil characteristics
of the calcareous marl are significantly different. The calcareous marl is stiffer than the black
marl and the grey marl. For the sake of simplicity, only two types of soil will be considered, for
the upper one and the lower one, we will use the same parameters. Convergence measurement
devices had been installed by CETU (Centre d’Etudes des Tunnel) during construction processes
in the tunnel at the sections PM 1168, PM 1156 and PM 1150. These sections, particularly for
the convergence D–E and F–G, have been selected to provide the simulation data in the
deformation analysis.

Figure 9. Dimensions, measurement positions and soil stratigraphy of the tunnel: (a) dimensions and
measurement positions; and (b) soil stratigraphy (after Lunardi [21]).

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
62 A. PURWODIHARDJO AND B. CAMBOU

Figure 9 shows the dimensions of the tunnel, the measurement positions (D–E and F–G) of
the tunnel and the soil stratigraphy at section PM 1168.

4.2. Model parameters


Elastoplastic and viscoplastic model parameters identified in this analysis can be seen in
Table III. Figure 10 shows the simulation results of triaxial tests by using elastoplastic soil
parameters. Figure 11 shows the simulation results of creep tests by using elasto-viscoplastic soil
parameters.
It can be seen from Figures 10 and 11 that the simulation results closely match the
experimental results. This means that this model can represent the elastoplastic and viscoplastic
behaviour of the soils quite satisfactorily.

4.3. Plane strain calculations


The convergence curves of the unsupported tunnel are derived by using virtual support
pressures in plane strain calculations [4,22].
Calculation of actual conditions would need a 3D analysis which would be very
time consuming for creep analysis. Therefore calculation has been performed using two

Table III. Parameters of the model.


Parameter Black and grey marls Calcareous marl
3
Density ðkN=m Þ 22.15 24.34
Elastoplastic parameters
G0 (MPa) 27 96.15
K0e (MPa) 139 208.33
K0p (MPa) 139 208.33
n 0.7 0.6
g 0.3616 0.7852
A ðkPa1 Þ 2 25
Rc peak ¼ Rcresidual 0.0784 0.213
b 0.005 0.38
m 0.033 0.05
c 60.0 75.0
Pc0 ðMPaÞ 17.00 40.00
m0 2.5 2.5
Tr peak ðMPaÞ 11.825 8.768
Tr residual ðMPaÞ 7.112 5.273
e0II 0.02 0.02
eIIf 0.065 0.065

Viscoplastic parameters
Av 125.0 450
Z0 ðsÞ 108 106:69
k 6.0 30.673
m1 0.3 0.4083
m2 0.0 8.0214

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
TIME-DEPENDENT MODELLING FOR SOILS 63

8 35

7 30

6
25

5
20

q (MPa)
q (MPa)

15
3
Exp. σ 3 = 2.60 MPa
Exp. σ3 = 2.25 MPa
Sim. σ 3 = 2.60 MPa 10
2 Sim. σ3 = 2.25 MPa
Exp. σ 3 = 4.12 MPa
Exp. σ3 = 5.00 MPa
Sim. σ 3 = 4.12 MPa 5
1 Sim. σ3 = 5.00 MPa

0 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0
(a) ε 1 (%) (b) ε 1 (%)

Figure 10. Numerical simulation results of triaxial tests (experiments from Serratrice [20]): (a) black and
grey marls; and (b) calcareous marls.

2.0 0.50

1.8 0.45 σο= 2.558 MPa

1.6 σο = 1.298 MPa 0.40

1.4 0.35

1.2 σο= 0.669 MPa 0.30 σο= 1.288 MPa


ε 1 (%)
ε 1 (%)

1.0 0.25

0.8 0.20

Exp. σ o = 0.669 MPa Exp. σo = 1.288 MPa


0.6 0.15
Sim. σ o = 0.669 MPa Sim. σo = 1.288 MPa
0.10 Exp. σo = 2.558 MPa
0.4 Exp. σ o = 1.298 MPa

Sim. σ o = 1.298 MPa Sim. σo = 2.558 MPa


0.2 0.05

0.0 0.00
0 100 200 300 400 500 600 700 800 900 1000 0 100 200 300 400 500 600 700 800 900
(a) Time (hours) (b) Time (hours)

Figure 11. Numerical simulation results of drained creep triaxial tests (experiments from Serratrice [20]):
(a) black and grey marls; and (b) calcareous marls.

kinds of calculation: a plane strain calculation to take into account the actual stress and shape
of the tunnel and an axisymmetric calculation to take into account the distance to the
working face.

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
64 A. PURWODIHARDJO AND B. CAMBOU

Two types of calculation have been accomplished. The first one is by using a circular shape
of tunnel with R ¼ 7:65 m; Ko ¼ 1:0 and only using one soil layer (black and grey
marl parameters), and the second one is by using the actual shape of the tunnel, the
actual value of Ko ¼ 1:5 and the actual soil layers. The objective of this calculation is to
provide a comparison of the results from those two shapes for the reason that the calculation
taking into account the distance from the working face and the progress of the tunnelling will be
performed using an axisymmetric calculation (see Section 4.4). The result comparison obtained
in the plane strain calculation allows the results obtained in the axisymmetric condition to be
transformed to take into account the actual conditions of the tunnel section. This
approximation has been made because those two shapes are almost similar and the calcareous
marl is not dominant, however it can be noted that this kind of calculation can only be
considered as a first approximation.
In the first calculation, a quarter of the tunnel geometry has been modelled, and for the soil
parameters, the black and grey marl parameters have been employed. On the other hand, in the
second calculation, both types of soil parameters have been employed. The overburden pressure
height is 100 m from the crown of the tunnel and the ground water table is 6:6 m below the
ground level.
Figure 12 shows the mesh used in the plane strain calculations and Figure 13 shows the
comparison of results of the convergence analysis of the two shapes. By comparing those two
results, we can determine the shape ratio of the unsupported tunnel which is defined by
convergenceactual shape
RT ¼ ð26Þ
convergencecircular shape
where RT is the shape ratio of the tunnel for a given monitoring location.
Figure 13 shows the comparison of convergence analysis of the unsupported tunnel. From
that figure, we can determine that the shape ratio of the tunnel for the convergence at positions

Figure 12. Mesh used in the plane strain calculations: (a) mesh used in the circular
shape; and (b) mesh used in the actual shape.

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
TIME-DEPENDENT MODELLING FOR SOILS 65

D–E and F–G equals 0.99. This ratio will be used for adjusting the axisymmetric calculations,
since in the axisymmetric calculations, we can only use the circular shape, Ko ¼ 1:0 and one
layer of soil. In this calculation, we suppose that the convergence ratio for those shapes of tunnel
(circular and actual shape) is similar for the supported and the unsupported tunnel.

4.4. Axisymmetric calculation


This computation is performed by using the sequential excavation method (SEM)
in the axisymmetric condition. Distance to the working face is defined by d and the
advance rate of the excavation is defined by p (see Figure 14). Figure 15 shows the geometry of

10

20

30
Actual shape - conv. F-G, Ko = 1.5

40
Circular shape - conv. F-G
λ (%)

50

60

70

80

90 Circular shape - conv. D-E


Actual shape - conv. D-E, Ko = 1.5
100
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
Ur (m)

Figure 13. Comparison of convergence analysis of the unsupported tunnel.

Figure 14. Sequential excavation method (SEM) in the axisymmetric calculations.

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
66 A. PURWODIHARDJO AND B. CAMBOU

Figure 15. Geometry of the tunnel in the axisymmetric calculations.

Figure 16. Mesh used in the axisymmetric calculations and measurement positions: (a) mesh used in the
axisymmetric calculations; and (b) zoom of mesh and measurement positions.

the tunnel in the axisymmetric calculation and Figure 16 shows the mesh used in the
axisymmetric calculation.
The lining support in this tunnel is a combination of shotcrete ring and steel frame. For
simplifying the analysis, the equivalent stiffness of the lining of the combined lining support has
been used. A linear elastic model is used for this lining.

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
TIME-DEPENDENT MODELLING FOR SOILS 67

For modelling the ground anchor on the working face, the equivalent pressure on the working
face has been used [23]. This pressure is determined by
 
nb Tb nb Sb
sf ¼ min ; ð27Þ
S S
where nb is the number of anchors, Tb is the tensile strength of the anchor, Sb is the shear
strength of the anchor, and S is the working surface area. In this tunnel, 120 fibre–glass anchors
with 800 kN tensile strength have been installed to stabilize the working face.
In this calculation, the black and grey marl parameters have been used; because, they are
more dominant than the calcareous marl in the soil stratigraphy. For the SEM analysis, we use
d ¼ 2:0 m; p ¼ 1:0 m (the advance length of tunnelling) and R ¼ 7:65 m (see Figure 9).
The convergence of the tunnel, which can be measured, is determined by
CðxÞ ¼ 2RT ðUðxÞ  UðoÞÞ ð28Þ

where UðxÞ is the deformation of the tunnel as a function of the distance from the working face,
UðoÞ is the deformation of the tunnel at the working face and RT is the shape ratio of the tunnel.
In the first simulation, the influence of the advance rate of the tunnelling is illustrated. Three
types of advance rate are used, i.e. 3.0, 1.5 and 0:75 m/day. The elastoplastic calculation is used
to represent the infinite advance rate of tunnelling. The tunnelling simulation results can be seen
in Figure 17.
From that figure, we can see that the convergence of the tunnel when the lining is applied can
be reduced by increasing the advance rate of tunnelling. If we only use the elastoplastic
constitutive model, we cannot illustrate this phenomenon. The total elasto-viscoplastic
deformation of the tunnel in this case is significantly bigger than the elastoplastic deformation
(instantaneous deformation).

0.50
3.0 m / day ; d = 2 m

0.45 1.5 m / day ; d = 2 m

0.75 m / day ; d = 2 m
0.40
Elastoplastic calculation
p = 1 m, d = 2m
0.35

0.30
Ur/R (%)

0.25

0.20

0.15

0.10

0.05

0.00
0 1 2 3 4
DISTANCE FROM WORKING FACE (x R)

Figure 17. Tunnelling simulation results from the elastoplastic–viscoplastic calculations.

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
68 A. PURWODIHARDJO AND B. CAMBOU

In the second simulation, the actual advance rate of tunnelling at PM 1156 and PM 1150 is
used. The sequences of the actual excavation are as follows:
(a) excavating the tunnel with the advance rate: 2:0 m/day until 12 m;
(b) stopping for 2 days (to represent the excavation of the lower section and the installation
time of the anchors on the working face),
(c) continuing the two sequences above.
Figure 18 presents the advance rate of tunnelling used in the numerical simulation.
The measurements at PM 1168 (see Figure 19) are started 6 m behind the working face while
for PM 1156 and PM 1150 are started 9 and 5 m behind the working face, respectively. Figure
19 shows the comparison of results from the calculations and the measurements, started from
the first measurement. This comparison is shown with respect both to the distance to the
working face and time. The evolution of the convergence linked to the creep of soil when the
excavation is stopped can clearly be seen both on the measurements and on the numerical
simulation. These two comparisons show the ability of the numerical modelling to represent well
the essential phenomena occurring during the tunnelling process.
Strain gauge devises were installed by CETU on the steel frame (HEB 300) at section PM
1168. By considering the elastic modulus of steel equal to 200 GPa; the measurement results can
be analysed in terms of the stress. The stress of the lining at the crown and the spring line was
365 MPa and the stabilized stress was achieved 30 days after the installation of the lining.
Figure 20 shows the numerical simulation of the evolution of the compression stress in the
lining. It can be seen that the stabilized stress was achieved twenty days after installation while
30 days for the measurements. If the stress measurement result is transformed in terms of the
equivalent stiffness of the combined lining supports we get 18:25 MPa while 21 MPa for the

55 1114

50 1119

45 1124
Progression of tunnelling (m)

40 1129

35 1134

30 1139
PM

25 1144

20 1149

15 1154

10 1159

5 1164

0 1169
0 5 10 15 20 25 30 35
Elapsed time of excavation (days)

Figure 18. Advance rate of tunnelling in the numerical simulation.

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
TIME-DEPENDENT MODELLING FOR SOILS 69

25 25

Simulation
CONVERGENCE D-E (mm) 20 Measurement 20

CONVERGENCE D-E (mm)


D Conv. D-E E
F Conv. F-G G
15 15

D Conv. D-E E
10 10
F Conv. F-G G

5 5 Simulation
Measurement

0 0
0 5 10 15 20 25 30 35 0 5 10 15 20
(a) DISTANCE FROM WORKING FACE (m) Time (days)

25 25

D Conv. D-E E
20 20
F Conv. F-G G
D Conv. D-E E
CONVERGENCE F-G (mm)

F Conv. F-G G
CONVERGENCE F-G (mm)

15 15
Simulation
Measurement
Simulation
10 Measurement 10

5 5

0 0
0 5 10 15 20 25 30 35 0 5 10 15 20
(b) DISTANCE FROM WORKING FACE (m) Time (days)

25 25

20 20
D Conv. D-E E
CONVERGENCE F-G (mm)
CONVERGENCE F-G (mm)

F Conv. F-G G

15 15
Simulation
Measurement

10 10 D Conv. D-E E
F Conv. F-G G

5 5 Simulation
Measurement

0 0
0 5 10 15 20 25 0 5 10 15 20
(c) DISTANCE FROM WORKING FACE (m) Time (days)

Figure 19. Comparison between numerical simulations and measurements: (a) convergence D–E at section
PM-1168; (b) convergence F–G at section PM-1156; and (c) convergence F–G at section PM-1150.

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
70 A. PURWODIHARDJO AND B. CAMBOU

Figure 20. Evolution of the compression stress in the lining.

calculation result (characteristic strength of concrete used is 25 MPa). So the numerical


calculation gives results in agreement with the measurement results.

5. CONCLUSIONS

A time-dependent model based on the framework of elastoplasticity–viscoplasticity has been


developed. The viscous mechanism in this model is governed by a particular hardening called
‘viscous hardening’ with a bounding surface. The validation results using classical laboratory
tests presented in this study show that this model is able to predict the undrained creep process
including accelerated creep and creep rupture. This model is able to describe the undrained
relaxation process.
The analysis of deformations due to tunnelling using the elastoplastic–viscoplastic
constitutive model showed that the influence of viscous effects cannot be neglected in the soil
which has been analysed. The initial deformation when the lining is applied can be reduced by
increasing the advance rate of tunnelling but attention should be paid to the lining because the
load transfer to the lining will be higher. This becomes significant when there is a large distance
between the installation point of the lining and the working face, and could induce plastic
deformation around the tunnel. The evolution of the compression stress in the lining can be
modelled quite satisfactory using the model proposed.
The calculation procedure proposed has provided an effective approach for analysing the
ground-structure interaction situation and offers a systematic way of optimizing lining design
and predicting accurately the evolution of the deformations and the stress around the tunnel.
This kind of calculation can be improved by using a complete 3D approach. However, this is a
rather difficult calculation and the computation time will be long. In practice, the plane strain
and the axisymmetric analysis can be successfully used to develop a reasonable solution.

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71
TIME-DEPENDENT MODELLING FOR SOILS 71

ACKNOWLEDGEMENTS

The authors gratefully acknowledge the information on the geotechnical data and the convergence data
provided by Mr. Alain Robert and Mr. Adrien Sa.ıtta from CETU (Centre d’Etudes des Tunnels), Lyon,
France.

REFERENCES
1. Cambou B, Jafari K. A constitutive model for granular materials based on two plasticity mechanisms. In
Constitutive Equations for Granular Non-Cohesive Soils, Saada A.S, Bianchini G (eds). Balkema: Rotterdam, 1987;
149–167.
2. Maleki M. Mod!elisation hi!erarchis!ee du comportement des sols. Ph.D. Thesis, Ecole Centrale de Lyon, Montreux, 1998.
3. Nguyen Minh D, Berest P. Etude de la stabilit!e des cavit!es souterraines avec un mod"ele de comportement
e! lastoplastique radoucissant. 4eme Cong. de la Soc. Int. De Me!c. Des Roches, London, 1979.
4. Panet M, Guenot A. Analysis of convergence behind the face of a tunnel. Tunnelling’82, 1982; 197–204.
5. Borosetto M, Ribacchi R. Influence of the strain-softening behaviour of rock masses on the stability of a tunnel.
Third International Conference on Numerical Methods in Geomechanics, Aachen, 1979; 611–620.
6. Brown ET, Bray JW, Ladanyi B, Hoek E. Ground response curves for rock tunnels. Journal of Geotechnical
Engineering Division, ASCE 1983; 109(1):15–39.
7. Aydan O . , Akagi T, Kawamoto T. The squeezing potentials of rocks around tunnels. Theory and prediction. Rock
Mechanics and Rock Engineering 1993; 26(2):137–163.
8. Perzyna P. Fundamental problems in viscoplasticity. Advances in Applied Mechanics 1966; 9:243–277.
9. Katona MG. Evaluation of viscoplastic cap model. Journal of Geotechnical Engineering 1985; 110(8):1107–1125.
10. Adachi T, Oka F. Constitutive equations for normally consolidated clays based on elasto-viscoplasticity. Soils and
Foundations 1982; 22(4):57–70.
11. Adachi T, Oka F, Mimura M. An elasto-viscoplastic theory for clay failure. Proceedings of the 8th Asian Regional
Conference on Soil Mechanics and Foundation Engineering, Kyoto, 1987; 5–8.
12. Adachi T, Oka F, Mimura M. Elasto-viscoplastic constitutive equations and its application to consolidation
analysis. Journal of Engineering Materials and Technology 1990; 112:202–209.
13. Sekiguchi H. Theory of undrained creep rupture of normally consolidated clays based on elasto-viscoplasticity. Soils
and Foundations 1984; 24(1):129–147.
14. Kaliakin N, Dafalias F. Theoretical aspects of the elastoplastic–viscoplastic bounding surface model for cohesive
soils. Soils and Foundations 1990; 30(3):11–24.
15. Akai K, Adachi T, Nishi K. Mechanical properties of soft rocks. Ninth International Conference on Soil Mechanics
and Foundation Engineering, Tokyo, 1977; 7–10.
16. Lade PV, Liu Chi-Tseng. Experimental study of drained creep behavior of sand. Journal of Engineering Mechanics
1998; 124(8):912–920.
17. Purwodihardjo A. Mod!elisation des d!eformation diff!er!ees lors du creusement des tunnels. Ph.D. Thesis, Ecole
Centrale de Lyon, 2004.
18. Piepi GT. Comportement viscoplastique Avec Rupture Des Argiles Raides. Ph.D. Thesis, Ecole Nationale des Ponts
et Chauss!ees, 1995.
19. Rousset G. Comportement m!ecanique des argiles et marnes profondes: application la mod!elisation des ouvrages
souterrains. Rapport Conseil Scientifique G3S 1989; 213–219.
20. Serratrice JF. Tunnel de Tartaiguille (Dro# me) TGV M!editerran!ee, Essais de laboratoire sur la marne, LRPC d’Aix
en Provence, 1999.
21. Lunardi P. The design and construction of tunnels using the approach based on the analysis of controlled
deformation in rocks and soils. ADECO-RS, 2000.
22. AFTES. La m!ethode convergence-confinement. Tunnels et Ouvrages Souterrains 2002; 170:79–89.
23. Peila D. A theoretical study of reinforcement influence on the stability of a tunnel face. Geotechnical and Geological
Engineering 1994; 12:145–168.

Copyright # 2004 John Wiley & Sons, Ltd. Int. J. Numer. Anal. Meth. Geomech. 2005; 29:49–71

You might also like