Vous êtes sur la page 1sur 14

Soil Dynamics and Earthquake Engineering 66 (2014) 206–219

Contents lists available at ScienceDirect

Soil Dynamics and Earthquake Engineering


journal homepage: www.elsevier.com/locate/soildyn

Numerical validation of analytical solutions and their use


for equivalent-linear seismic analysis of circular tunnels
S. Kontoe n, V. Avgerinos, D.M. Potts
Department of Civil & Environmental Engineering, Imperial College London, South Kensington Campus, London SW7 2AZ, UK

art ic l e i nf o a b s t r a c t

Article history: The first part of this paper presents an extensive validation of four analytical solutions for the seismic
Received 12 December 2013 design of circular tunnels. The validation is performed with a quasi-static finite element (FE) model
Received in revised form which conforms to the assumptions of the analytical solutions. Analyses are performed for a wide range
20 June 2014
of flexibility ratios, slippage conditions at soil–lining interface, assuming both drained and undrained
Accepted 8 July 2014
behaviour. Based on the numerical predictions the relative merits of the considered analytical solutions
are discussed and recommendations are given for their use in design. The second part of this paper
Keywords: explores the use of equivalent linear soil properties in analytical solutions as an approximate way of
Quasi-static finite element analysis simulating nonlinearity. The results of equivalent linear site response analyses are used as an input for
Seismic design of tunnels
the analytical solutions. The comparison of the analytical predictions with nonlinear numerical analysis
Equivalent linear site response analysis
results is very satisfactory. The results of this study suggest that analytical solutions can be used for
preliminary design using equivalent linear properties and the corresponding compatible strain as an
approximate way of accounting for nonlinear soil response.
& 2014 Elsevier Ltd. All rights reserved.

1. Introduction degree of slippage and consequently it is difficult to select an


appropriate value for the flexibility coefficient.
The seismic design of circular tunnels is currently based on The most widely used and discussed analytical solutions are the
analytical solutions [11,19,4,5,18], deformation based quasi-static generalised method of Hoeg [11] (often referred to as Wang [25])
finite element (FE) analysis [15,10,22,2]and time-domain FE ana- and the method of Penzien [19]. These two approaches result in
lysis [15,14,1,6,2,16] which simulate vertical propagation of shear identical seismic loads for the case of full-slip assumption along the
waves. The above mentioned analytical solutions and quasi-static soil–structure interface, whereas for the no-slip case significant
FE analysis consider the lining ground interaction, but they neglect discrepancies are observed [9,15,18]. Hashash et al. [10] and more
any inertial interaction which can only be examined with a time- recently Sedarat et al. [22] compared the results of simple quasi-
domain analysis. static finite element analyses with the two analytical solutions in
The analytical solutions are very attractive tools for preliminary terms of axial force and bending moments for the no-slip case. These
design, as they provide a quick and easy calculation of the seismic studies not only confirmed the differences between the two con-
design loads in the tunnel lining in terms of axial force and bending sidered analytical solutions, but they also revealed that the Penzien
moment. Most analytical solutions assume either zero friction (full- [19] approach severely underestimates the seismic axial forces.
slip condition) or full connexion (no-slip condition) between the soil Another important assumption which has been adopted by
and the lining. Park et al. [18] provides the only analytical solution, most analytical solutions is that of dry soil conditions. Naturally,
to the authors' knowledge, that considers the slippage effect at this assumption is not realistic when tunnels are placed in
the soil–lining interface using a spring-type flexibility coefficient. saturated deposits and are subjected to rapid loading. Bobet [4,5]
As suggested by many studies [9,22,16] for most cases the interface dealt with this limitation and provided a solution for undrained
condition is actually somewhere between full-slip and no-slip and, conditions.
in that respect, the contribution of the Park et al. [18] solution is In the first part of this paper, an extended validation of four
significant. In practice though, it is difficult to, a-priori, quantify the analytical solutions [11,19,4,5,18] is presented for various slippage
conditions, a wide range of flexibility ratios and for both drained
and undrained conditions. The validation is performed using linear
n
Corresponding author. Tel.: þ 44 20 7594 5996; fax: þ 44 20 7594 5934.
elastic quasi-static analyses with the finite element code ICFEP [20].
E-mail addresses: stavroula.kontoe@imperial.ac.uk (S. Kontoe), The second part of this paper explores the use of equivalent
v.avgerinos@imperial.ac.uk (V. Avgerinos), d.potts@imperial.ac.uk (D.M. Potts). linear properties in analytical solutions as an approximate way of

http://dx.doi.org/10.1016/j.soildyn.2014.07.004
0267-7261/& 2014 Elsevier Ltd. All rights reserved.
S. Kontoe et al. / Soil Dynamics and Earthquake Engineering 66 (2014) 206–219 207

simulating soil's nonlinearity, which was previously suggested by accepted though that two-dimensional plane strain models pro-
Wang [25] and later employed by Hashash et al. [9] and Amorosi and vide a reasonable approximation of the problem, as the most
Boldini [1]. A recent study by Argyroudis and Pitilakis [2] shows a critical mode is the ovaling deformation of the tunnel's cross
poor comparison between analytical solutions and nonlinear quasi- section [19] which is schematically illustrated in Fig. 1a.
static FE analysis, thus suggesting the exclusive use of fragility curves In the present study four analytical solutions (generalised
derived by numerical analysis for preliminary design. This study method of Hoeg's [11] (often referred to as Wang [25]), Penzien
shows that the results of most analytical solutions when adopting [19], Park et al. [18] and Bobet [5]) are examined and compared
equivalent linear properties compare very favourably with nonlinear with numerical results, concerning the bending moments, axial
FE analysis and provides possible explanations for the discrepancy forces and lining deflection due to ovaling deformation. When
with the Argyroudis and Pitilakis [2] findings. employing such analytical solutions it is very important to first
appreciate their underlying assumptions:

2. Problem statement and analytical solutions  The tunnel lining (for the seismic loading) is considered to act as
an elastic beam under a uniform shear-strain field of amplitude
It is widely accepted [17] that the response of circular tunnels γmax, ignoring any inertial soil–lining interaction.
to ground shaking can be described by three main deformation  The initial stresses in the ground and the lining are not considered
modes: (a) axial deformation along the tunnel, (b) longitudinal and the excavation is assumed to take place simultaneously with
bending and (c) ovaling deformation of the tunnel's cross section. the construction of the lining.
The first two deformation modes occur due to seismic waves  The strain is applied under fully drained conditions (note that
propagating parallel or at an angle to the tunnel axis, while the Bobet [5] gives a separate solution for undrained conditions).
latter one is related to seismic shear waves propagating perpendi-
cularly to the tunnel axis.
Obviously, in order to simulate rigorously all these three modes It should be mentioned that the dynamic interaction can be of
of deformation, a three-dimensional model is needed. It is widely importance when (a) the dimensions of the tunnel's cross section

Fig. 1. (a) Ovaling deformation due to vertically propagating shear waves (adapted from [17]), (b) corresponding seismic shear loading and (c) equivalent static loading
(adapted from [18]).
208 S. Kontoe et al. / Soil Dynamics and Earthquake Engineering 66 (2014) 206–219

are comparable to the wavelengths of the predominant frequency


of the earthquake loading, (b) the tunnel is relatively shallow and
(c) for stiff structures in soft soil. The loading in the tunnel under
static conditions depends on the stiffness of the lining, the over-
burden pressure and the coefficient of earth pressure at rest, K0.
For the seismic case, the free-field shear stress replaces the in-situ
overburden pressure and the at-rest coefficient of earth pressure is
assigned a value of (  1). Hence the loading induced by shear
waves is assumed as being equivalent to a compressive and tensile Fig. 2. Schematic representation of the mesh configuration in a quasi-static
principal stress acting at 451 with respect to the horizontal axis (as deformation based analysis consistent with simple-shear conditions.
shown in Fig. 1b and c). Clearly due to this assumption, regarding
the seismically induced stress state in the field, all these analytical Table 1
solutions predict maximum ovaling deformation at the major and Soil parameters and corresponding flexibility ratios.
minor axes at θ¼ 7 451 from the horizontal axis.
The main parameters that govern the predicted loads by the Soil's Young's modulus Εm Flexibility ratio
(kPa) F
analytical solutions are the compressibility, C, and the flexibility, F,
ratios and the shear-strain γmax. The shear strain is calculated 2018.2 0.12
either as the ratio of the particle velocity to the shear wave 7427.1 0.46
propagation velocity or, preferably, from a site response analysis 16,145.8 1.00
as the maximum shear strain at the level of the tunnel axis. 35,843.7 2.22
350,455.7 21.7
The compressibility, C, and the flexibility, F, ratios are means to 1,120,000.0 69.4
quantify the relative stiffness of the medium compared to that of 2,240,000.0 138.7
the lining. The compressibility ratio is a measure of the extensional 3,000,000.0 185.8
stiffness of the medium relative to that of the lining and the
flexibility ratio is a measure of the flexural stiffness
The assumed model represents a soil layer of thickness of
Em ð1  ν2l Þr H¼ 50 m overlying bedrock. The vertical boundaries of the mesh
C¼ ð1Þ
El tð1 þ νm Þð1  2νm Þ are 50 m away from the centre of the tunnel. The soil was con-
sidered to behave elastically with Poisson's ratio νm ¼0.25, while its
Em ð1 ν2l Þr 3 Young's modulus, Em , was varied parametrically in order to capture
F¼ ð2Þ
6El Ið1 þ νm Þ a wide range of flexibility ratios F (Table 1). The modelled tunnel has
a circular cross section of radius r¼ 3 m and its centre is located
where Em and El are Young's moduli of the medium and the lining
15 m below the ground surface. The lining's width is t¼0.3 m and
respectively, νm and νl are Poisson's ratios of the medium and the
was modelled with elastic beam elements of Young's modulus
lining respectively, r and t are the radius and the thickness of the
El ¼24.8 GPa and Poisson's ratio, νl ¼0.2. Prior to all quasi-static
lining, respectively and I is the moment of inertia per unit width of
analyses, a static analysis was undertaken to establish the initial
the lining. Values of the flexibility ratio F greater than 1 suggest
stresses acting on the lining. During the static analysis, displace-
that the tunnel lining has lower stiffness than the surrounding
ments were restricted in both directions along the bottom mesh
medium. Hence, values of F-1 imply that the tunnel undergoes
boundary and horizontal displacements were restricted along the
identical deformations to that of an unlined tunnel without
side boundaries. The tunnel construction was modelled using the
resisting the ovaling deformation. The analytical solutions exam-
convergence-confinement method which is described in detail by
ine separately the cases of full and no-slip condition along the
Potts and Zdravković [21]. The tunnel excavation was performed in
interface of ground and lining. Slip at the interface is generally
10 increments and the linings were constructed at 50% of stress
possible depending on the construction method, the relative
relaxation (i.e. increment 5) prior to the completion of excavation.
stiffness at the soil–lining interface and the severity of the seismic
In order to be consistent with the assumptions of the analytical
loading, while as many studies suggest [9,22,16] in most cases the
solutions (which ignore the initial stresses in the ground and the
reality is somewhere in between full-slip and no-slip. For com-
lining), in all cases the “seismic increment” was obtained by
pleteness the four analytical solutions are presented concisely in
subtracting the loads computed at the end of the excavation from
the Appendix.
those at the end of the quasi-static analysis.
For the modelling of the slippage, interface elements of zero
thickness were used [8] with a very high normal stiffness (KN ¼
3. Numerical validation of analytical solutions 107 kN/m3) and varying shear stiffness depending on the degree of
slippage. In particular for the full-slip case a nominal value of
3.1. Numerical model and parameters considered shear stiffness (Ks ¼1 kN/m3) was used. In order to be consistent
with the assumptions of the analytical solutions, the vertical and
The validation of the analytical solutions was based on a simple horizontal displacements were restricted along the bottom bound-
quasi-static model which was created with the finite element code ary, while the vertical displacement was also restricted along the
ICFEP [20]. Quasi-static finite element analysis can either be forced side and top boundaries of the model. Furthermore, a uniform
based, in which seismically induced inertia forces are approxi- displacement u was applied along the top boundary, while a
mated as a constant body force, or deformation based, in which triangular displacement distribution was applied along the vertical
the mesh is subjected to shear deformation, as schematically boundaries of the mesh (see Fig. 2). The magnitude of the uniform
illustrated in Fig. 2. In this study a deformation based quasi- applied displacement u was u ¼0.126 m corresponding to a shear
static methodology is adopted, as this approach is consistent with strain:
the assumptions of the analytical solutions and it does not suffer
from the inherent limitations of the forced based approach (e.g. u 0:126 m
γ max ¼ ¼ ¼ 0:00252 ð3Þ
sensitivity on the dimensions of the model) [13]. H 50 m
S. Kontoe et al. / Soil Dynamics and Earthquake Engineering 66 (2014) 206–219 209

Fig. 3. Comparison of calculated thrust (a) and bending moments (b) by analytical solutions and numerical analysis for the full-slip condition.

This value of shear strain is identical to one adopted by Hashash which corresponds to Ks values of 106–104 kN/m3 for the shear
et al. [9] in their validation study. stiffness of the interface elements. The numerical results compare
very well with the analytical solution both in terms of thrust and
3.2. Drained analysis bending moments for all the considered flexibility coefficients.
The flexibility coefficient D has a more pronounced effect on the
In this section, the results of a series of numerical analyses for resulting thrust values rather than the bending moments with the
both full-slip and no-slip conditions are compared with the four effect becoming progressively larger as the flexibility ratio (F)
analytical solutions. Fig. 3 presents the comparison for the full-slip increases.
assumption in terms of lining thrust and bending moment. The Figs. 5 and 6 present the comparison for the no-slip case again
resulting thrust and bending moment for all analytical methods in terms of lining thrust and bending moment. In Fig. 5 the results
coincide, while the analytical results are also in excellent agree- for the lower examined flexibility ratios are shown (up to F ¼2.22),
ment with the numerical results for all examined flexibility ratios. while Fig. 6 presents the results for the higher flexibility ratios.
It should be noted that Sedarat et al. [22] performed a similar Concerning the thrusts, Hoeg, Park et al. and Bobet solutions
comparison for three flexibility ratios and two lining thickness match perfectly the numerical solution for the lower flexibility
(t ¼0.36 m and t¼1.26 m). They concluded that the results of ratios (Fig. 5a), while the comparison is also very good for the
extended Hoeg and Penzien for the full-slip condition in terms larger ones (Fig. 6a). However Penzien's solution gives comparable
of thrust loads are comparable with the numerical results only for results with the other approaches only for very low flexibility
the thick lining and for low values of the flexibility ratio (up to ratios, while as the flexibility ratio increases the Penzien method
3.2). In the present study although the lining thickness (t¼ 0.3 m) results diverge from the numerical solution considerably, reaching
is similar to the thin lining case of Sedarat et al. [22] (which differences of 28.3% and 46.3% for flexibility ratios F ¼1.0 and
showed poor agreement with the analytical solutions) the com- F ¼2.22 respectively. The underestimation of the thrust values by
puted thrust values herein are in excellent agreement with the this approach becomes even worse for greater flexibility ratios, as
analytical results for all examined flexibility ratios. It should be the predicted thrust values are lower by at least an order of
noted that Sedarat et al. [22] modelled the full-slip case with a magnitude compared to the other approaches and the numerical
frictionless contact, adopting an elastoplastic law for the soil– solution. This is in agreement with the findings of Hashash et al.
lining interface. This prohibits the development of tensile normal [10] and Sedarat et al. [22].
contact stresses and leads to the observed discrepancies, for the With regards to the bending moments, the comparison between
full-slip case, with the numerical results of the present study and Hoeg's or Penzien's approaches and the numerical solution is very
the analytical solutions. While the modelling of the interface good only for very low flexibility ratios F o1.0, while the results
condition in the present study is consistent with the assumptions of Park et al.'s and Bobet's solutions are matching the numerical
of analytical solutions, the use of an elastoplastic law for the ones for even larger flexibility ratios (up to F ¼2.22 in Fig. 5b). For
interface condition is more realistic providing that it is accom- flexibility ratios F 421.7 (Fig. 6b) the numerical solution predicts a
panied by an elastoplastic model for the surrounding soil. very small change in bending moments, while the analytical
Among the existing analytical solutions, Park et al. [18] is the solutions result in an increase of its value with increasing flexibility
only one that considers the slippage effect at the soil–lining ratio (with a reduced rate though). The percentage differences of
interface employing a spring-type flexibility coefficient (D). the bending moments between analytical and numerical solutions
According to Park et al. [18], while the flexibility coefficient can are summarised in Table 2. The superiority of the Park et al. [18] and
theoretically vary from D ¼0 (corresponding to the no-slip case) to Bobet [5] solutions is clear as the maximum observed deviation,
D-1 (representing the full-slip case), in practice it is limited to a for the flexibility ratios examined herein, is almost half of the other
narrower range. Fig. 4 compares the calculated thrust and bending two approaches (15% compared to 27% deviation of the other two
moments with the Park et al. [18] solution for D ¼10  6–10  4 solutions).
210 S. Kontoe et al. / Soil Dynamics and Earthquake Engineering 66 (2014) 206–219

Fig. 4. Comparison of calculated thrust (a) and bending moments (b) by Park et al. [18] and numerical analysis for various slippage levels.

Fig. 5. Comparison of calculated thrust (a) and bending Moments (b) by analytical solutions and numerical analysis for the no-slip condition and for F¼ 0.125, 0.46,
1.0 and 2.22.

In order to compare the thrusts and moments (for the no-slip Furthermore Fig. 8 compares the results of the analytical and
assumption) of the analytical procedures with the numerical numerical solutions in terms of normalised lining deflection at
solution for a greater range of flexibility ratios, all the above θ¼ 7451 from the horizontal axis, which is defined as follows:
calculations and analyses were repeated for an increased tunnel
Δdl Δdl
radius of r ¼5 m. All the remaining parameters concerning both ¼7 ð4Þ
Δdfreefield rγ max
the soil and the lining were kept the same. The larger radius led to
an increase of all the examined flexibility ratios. The results are where Δdl is the lining deflection, Δdf reef ield is the corresponding
summarised in Fig. 7, while a logarithmic axis is used because the diametric strain for a circular cross section for non-perforated
range of the investigated flexibility ratios varied from less than ground, r is the tunnel radius and γmax is the maximum free-field
1 to almost 650. The results, both for thrusts (Fig. 7a) and for shear-strain at the level of the tunnel axis. In both cases (i.e. full-
bending moments (Fig. 7b), confirm the observations of Figs. 5 slip and no-slip) the discrepancy between the numerical predic-
and 6 and show that the discrepancy between analytical and tions and the analytical solutions is significant for flexibility ratios
numerical solutions in terms of bending moment increases with less or equal to one, while the comparison improves for higher
increasing flexibility ratio. values of flexibility ratio. For the full-slip case Hoeg's, Penzien's
S. Kontoe et al. / Soil Dynamics and Earthquake Engineering 66 (2014) 206–219 211

Fig. 6. Comparison of calculated thrust (a) and bending moments (b) by analytical solutions and numerical analysis for the no-slip condition and for F¼ 21.7, 69.4, 138.7
and 185.8.

Table 2
Percentage differences (%) between calculated maximum bending moments by full-slip case, showing a good agreement between the analytical
analytical solutions and numerical analysis for the no-slip assumption. and the numerical solution both in terms of thrust and bending
moments with the corresponding percentage differences reaching
F Extended Hoeg Penzien [19] Park et al [18] Bobet [5]
almost 6% for the thrust and 10% for the bending moment for the
0.12 11.93 0.19  1.28  1.28
high flexibility ratios (Table 3). The deviation between the numer-
0.46 13.63 3.86  0.62  0.62 ical and analytical solution is more significant for the no-slip case
1.00 15.64 7.97 0.34 0.34 as depicted in Fig. 10 and Table 3. The numerical analysis predicts
2.22 18.04 12.88 1.51 1.51 larger values of thrust and bending moment than Bobet [5] for low
21.7 22.89 22.06 5.18 5.18
flexibility ratios, while this trend is reversed for F421.7. While the
69.4 25.12 24.86 9.34 9.34
138.7 26.85 26.72 13.35 13.35 comparison in terms of bending moments is within acceptable
185.7 27.63 27.53 15.33 15.33 level the differences in terms of thrust values are significant.
It should be noted that the expressions of Bobet [5] for dry ground
and saturated ground under undrained conditions give identical
and Park et al.'s solutions give identical results, while the Bobet [5] results if a Poisson ratio, νm ¼ 0:5 is adopted.
solution gives lower values for the entire range of flexibility ratios.
For the no-slip case and for F 41 the solutions of Park et al. [18]
and Bobet [5] coincide and are in better agreement with the 4. Numerical evaluation of analytical solutions for nonlinear
numerical predictions than the remaining two solutions. It should response
be noted that Hashash et al. [10] observed a good agreement
between Hoeg's and Penzien's approaches and the numerical The use of the analytical solutions in seismic design usually
solution in terms of normalised lining deflection examining the involves the adoption of the maximum shear modulus and the
no-slip case (percentage differences up to 15.5%), but they con- maximum free-field shear-strain as input parameters. This can
sidered a limited range of flexibility ratios (F ¼9.63, 16.2 and 18.58) lead to erroneous estimation of the lining loads, as it is well esta-
within which the present study also shows a satisfactory compar- blished that soil behaves nonlinearly during seismic shaking and
ison (12% for Hoeg and 11% for Penzien for F ¼21.71). can experience significant stiffness degradation. To improve this
shortfall of the analytical solutions, it has been suggested [25] to
3.3. Undrained analysis perform first an equivalent-linear site response analysis of the
stratigraphy and to then use as input parameters for the analytical
Bobet [4] extended the analytical solutions for dry ground to solutions the equivalent linear (converged value) of shear modulus
fully saturated ground under undrained conditions, for both full and the corresponding “effective” shear-strain (which is a fraction
slip and no-slip conditions and expressed the bending moment of γ max ) at the level of the tunnel axis. In order to verify the validity
and the thrust as a function of the quasi-static free-field vertical of the suggested approach, the quasi-static finite element analyses
stress ðσ v Þ and horizontal stress ðσ h Þ. Later, Bobet [5] presented the were repeated adopting a nonlinear model using for the input
equations as a function of the far-field shear stress (as shown in deformation the effective shear-strain of each equivalent analysis
the Appendix) and further extended the methodology for the case at the depth of the tunnel axis.
of rectangular tunnels.
In order to validate the expressions proposed by Bobet [5] for 4.1. Equivalent linear analysis and nonlinear quasi-static analysis
undrained conditions, the set of analyses for a tunnel of radius
r ¼3.0m was repeated with the water table at the ground surface To estimate the equivalent linear shear stiffness and the
and undrained conditions. Fig. 9 presents the comparison for the compatible free-field shear-strain (γ ef f ) at the level of the tunnel
212 S. Kontoe et al. / Soil Dynamics and Earthquake Engineering 66 (2014) 206–219

Fig. 7. Comparison of calculated thrust (a) and bending moments (b) by analytical solutions and numerical analysis for the no-slip condition and for a wide range of
flexibility ratios (for tunnel with radius r¼ 5.0 m).

Fig. 8. Comparison of normalised lining deflection calculated by analytical solutions and numerical analysis for a wide range of flexibility ratios. (a) Full-slip, and (b) No-slip.

axis, site response analyses of a uniform one dimensional soil along with the results of the site response analysis in terms of
column of 50 m depth were performed with the software EERA equivalent shear stiffness and compatible free-field shear-strain
[3]. The column was subjected to the north–south component of (γ ef f ) at the level of the tunnel axis. It should be noted that the
the Griffith observatory motion recorded during the 1994 North- ratio of γ ef f to γ max was taken equal to 0.6 for all site response
ridge earthquake. This record was selected because of its very rich analyses, which according to the empirical expression of Idriss and
frequency content which is expected to engage a wide range of Sun [12] implies an earthquake of moment magnitude M ¼7.0.
modes of the problem. A parametric analysis was undertaken Subsequently, for all cases shown in Table 4, nonlinear elastic
varying the intensity of the ground motion and the soil type, quasi-static finite element analyses were carried out employing
aiming to cover a wide range of shear strains. In particular, the the boundary conditions and the lining parameters described
input motion was scaled to maximum peak ground accelerations previously for the linear case and assuming no-slip between the
of 0.2 g and 0.4 g and two soil types, a clean sand with plasticity soil and the lining. These nonlinear elastic analyses were carried
index PI¼0 and a medium plasticity clay with PI¼50, were out with a cyclic nonlinear model for the soil which adopts
examined. The adopted stiffness degradation and damping curves a hyperbolic function for the backbone curve as described in
were based on the expressions of Darendeli [7] and were [23, 24]. The model follows the Masing rules for unloading and
also subsequently used to calibrate the nonlinear model of the reloading, but this feature of the model was not used herein as
FE quasi-static analysis. Table 4 summarises the examined cases only monotonic loading was imposed for the purposes of this
S. Kontoe et al. / Soil Dynamics and Earthquake Engineering 66 (2014) 206–219 213

Fig. 9. Comparison of calculated thrust (a) and bending moments (b) by Bobet [5] and numerical analysis for undrained conditions and full-slip assumption.

Table 3
Percentage differences (%) between calculated maximum thrust and bending
flexibility ratio. The higher the flexibility ratio, the stiffer the soil
moments by Bobet [5] and numerical analysis for undrained conditions. and hence the lower the value of effective shear-strain that results
from the site response analysis (see Table 4). This explains the
F Full-Slip No-Slip differences observed in the trends of linear and nonlinear analyses,
while it indicates that using the elastic properties can result in
Tmax Mmax Tmax Mmax
overestimation of the lining loads for high flexibility ratios.
0.12  3.27  0.15  360.74  14.78
0.46  0.65 2.00  149.66  9.24 4.2. Comparison with analytical solutions
1.00 1.30 3.79  63.89  5.45
2.22 3.05 5.55  30.57  2.38
5.59 8.84  6.75 1.77
Figs. 12 and 13 compare the nonlinear analyses results with the
21.7
69.4 5.88 9.93 0.85 7.43 analytical solutions for PGA values of 0.2 g and 0.4 g respectively.
138.7 5.70 9.70 5.56 13.49 As previously explained Geq and the compatible free-field shear-
185.7 5.59 9.68 7.67 16.35 strain (γ ef f ) at the level of the tunnel axis were used as input
parameters for the analytical solutions. Similar to the linear case
(see Figs. 5 and 6) Park et al. [18] and Bobet [5] solutions compare
study. The model was calibrated to the stiffness degradation very well with the nonlinear FE analyses, Hoeg's solution shows
curves that were used for the site response starting with the larger differences in terms of bending moments, while Penzien
elastic shear stiffness (Gmax) values shown in Table 4. For each [19] underestimates the thrust values. The corresponding percen-
analysis the imposed deformation was calculated based on Eq. (3) tage differences between analytical solutions and numerical ana-
using the compatible free-field shear-strain (γ ef f ) at the level of the lyses are shown in Tables 5 and 6 for PGA values equal to 0.2 g and
tunnel axis shown in Table 4. In order to be consistent with the 0.4 g respectively. Overall the comparison between the analytical
assumptions of the analytical solutions, the simulation of the solutions and numerical analysis for the thrust values deteriorates
excavation and the tunnel construction was undertaken assuming with increasing nonlinearity (i.e. higher PGA level and lower
linear elastic behaviour (using Gmax) and the nonlinear model was flexibility ratios), but it is always within acceptable levels (exclud-
activated only for the quasi-static part of the analysis. In all cases ing Penzien for thrust predictions). Furthermore Fig. 14 shows the
the “seismic increment” was obtained by subtracting the loads comparison between numerical analysis and analytical solutions
computed at the end of the excavation from those at the end of the in terms of normalised lining deflection. Again the agreement
quasi-static analysis. between the two methodologies (i.e. analytical and numerical) is
Fig. 11 shows the thrust and bending moment variation with good with the Hoeg and Penzien solutions showing slightly better
flexibility ratio resulting from the nonlinear FE analyses for all comparison with the numerically obtained deflections. These
cases. It should be noted that the flexibility ratios have been observations confirm that analytical solutions with equivalent
calculated using the equivalent linear values of shear modulus linear shear stiffness and the corresponding compatible shear-
shown in Table 4. The thrust values seem to reach, with some strain at the tunnel axis level as input parameters can be success-
fluctuations, a plateau for high flexibility ratios, while for the fully used for the preliminary seismic design of circular tunnels.
linear case they were monotonically increasing with increasing The proposed approach can lead to more accurate estimation of
flexibility ratio (see Fig. 7a). On the other hand, the bending the lining loads as soil nonlinearity is approximated in a consistent
moments reduce with increasing flexibility ratio, while in the manner.
linear case they were reaching a plateau for the high flexibility Recently, Argyroudis and Pitilakis [2] performed a similar com-
ratio (see Fig. 7b). It should be noted though that, while all linear parison between analytical solutions and FE quasi-static analysis
elastic analyses were performed for the same value of shear-strain, observing a poor agreement. Therefore they concluded that
in the nonlinear case the imposed shear-strain value varies with numerically derived fragility curves should be used for preliminary
214 S. Kontoe et al. / Soil Dynamics and Earthquake Engineering 66 (2014) 206–219

Fig. 10. Comparison of calculated thrust (a) and bending moments (b) by Bobet [5] and numerical analysis for undrained conditions and no-slip assumption.

Table 4
Parameters considered in site response analyses and corresponding results in terms of equivalent shear modulus and effective shear strain at the level of the tunnel axis.

PI PGA (g) Gmax (kPa) Geq (kPa) F Geq/Gmax γ ef f (%)

0 0.2 140,182.3 72,811.27 11.27 0.52 0.0465


448,000.0 273,580.5 42.36 0.61 0.0310
896,000.0 657,803.7 101.85 0.73 0.0167
1,200,000.0 1,008,380.0 156.14 0.84 0.0083
50 0.2 14,337.48 8364.264 1.30 0.58 0.0850
140,182.3 95,935.35 14.85 0.68 0.0527
448,000.0 336,672.5 52.13 0.75 0.0367
896,000.0 812,796.8 125.85 0.91 0.0102
1,200,000.0 1,075,343.0 166.50 0.90 0.0117
0 0.4 140,182.3 38,412.16 5.95 0.27 0.1465
448,000.0 239,162.0 37.03 0.53 0.0437
896,000.0 559,027.3 86.56 0.62 0.0291
1,200,000.0 860,149.2 133.18 0.72 0.0183
50 0.4 140,182.3 79,797.1 12.36 0.57 0.0905
448,000 289,267.6 44.79 0.65 0.0637
896,000 748,747.8 115.94 0.84 0.0207
1,200,000 1,017,085.0 157.48 0.85 0.0182

Fig. 11. Thrust and bending moment variation with flexibility ratio calculated with nonlinear quasi-static FE analysis.
S. Kontoe et al. / Soil Dynamics and Earthquake Engineering 66 (2014) 206–219 215

Fig. 12. Comparison of calculated thrust (a) and bending moments (b) by analytical solutions and nonlinear numerical analysis for PGA ¼0.2 g (no-slip condition).

Fig. 13. Comparison of calculated thrust (a) and bending moments (b) by analytical solutions and nonlinear numerical analysis for PGA ¼0.4 g (no-slip condition).

Table 5
Percentage differences (%) between calculated maximum thrust and bending Table 6
moments by analytical solutions and nonlinear numerical analysis for PGA ¼ 0.2 g. Percentage differences (%) between calculated maximum thrust and bending
moments by analytical solutions and nonlinear numerical analysis for PGA ¼ 0.4 g.
PI F Extended Hoeg Penzien Park et al. Bobet
PI F Extended Hoeg Penzien Park et al. Bobet
Tmax Mmax Tmax Mmax Tmax Mmax Tmax Mmax
Tmax Mmax Tmax Mmax Tmax Mmax Tmax Mmax
0 11.27 15.47 17.01  352.15 15.42 15.70  2.01 15.70  2.01
42.36 13.11 22.31  1336.53 21.87 13.16 5.04 13.16 5.04 0 5.95 22.69 10.88  154.93 7.99 23.07  8.93 23.07  8.93
101.85 7.39 27.27  3115.64 27.09 7.39 12.88 7.39 12.88 37.03 16.99 18.63  1120.89 18.11 17.04 0.37 17.04 0.37
156.14 4.04 28.51  4548.19 28.39 4.03 15.71 4.03 15.71 86.56 12.13 23.12  2570.96 22.90 12.14 7.45 12.14 7.45
50 1.30 5.06 14.81  40.40 7.82 6.24  1.19 6.24  1.19 133.18 7.78 27.78  3858.79 27.65 7.77 14.32 7.77 14.32
14.85 9.43 21.13  507.30 19.94 9.62 2.98 9.62 2.98 50 12.36 14.48 18.46  392.64 17.01 14.70  0.27 14.70  0.27
52.13 7.66 25.86  1727.19 25.52 7.70 9.69 7.70 9.69 44.79 12.41 23.57  1419.93 23.16 12.46 6.67 12.46 6.67
125.85 2.87 28.30  3890.11 28.16 2.86 14.74 2.86 14.74 115.94 4.53 27.98  3576.62 27.83 4.53 14.11 4.53 14.11
166.50 3.40 28.74  4808.00 28.63 3.38 16.22 3.38 16.22 157.48 3.78 28.54  4590.82 28.43 3.77 15.79 3.77 15.79
216 S. Kontoe et al. / Soil Dynamics and Earthquake Engineering 66 (2014) 206–219

Fig. 14. Comparison of normalised lining deflection calculated by analytical solutions and numerical analysis for a wide range of flexibility ratios (no-slip condition).
(a) PGA ¼ 0.2g, (b) PGA ¼0.4g.

design instead of analytical solutions. They attributed the observed A further important reason for the discrepancy between the
discrepancies mainly to the following reasons: numerical results and the analytical solutions in [2] is the adopted
boundary conditions in their quasi-static model which are not in
 The use of an elastoplastic constitutive model for their FE anal- agreement with the racking boundary conditions which are
ysis which can lead to redistribution of stresses around the assumed by the analytical solutions. In particular, vertical displa-
tunnel lining and plastic deformations. cement was not restricted along the top boundary of their model
 The simulation of the excavation which essentially modifies the (as shown in Fig. 7a of [2]). Furthermore the variation of strain
initial stress conditions around the tunnel prior to the quasi- with depth which was obtained by the site response analysis was
static analysis. imposed to the quasi-static FE model instead of a uniform strain
 The application of a deformation variation, with depth while an corresponding to the level of the tunnel axis. A direct consequence
average value of strain at the tunnel axis is considered for the of their adopted boundary conditions is that the numerical model
analytical solutions. did not predict the maximum lining response at θ ¼ 7 45o to the
horizontal axis (shown in Fig. 7b of [2]).
In order to compare numerical analysis with the analytical
solutions, within the context of a validation exercise, is important
to model as closely as possible the assumptions of the analytical 5. Conclusions
solutions. Once the validation exercise is completed, a further
comparison can be carried out with more sophisticated numerical The first part of this paper examined parametrically the
modelling to assess the validity of the assumptions of the anal- reliability of four analytical solutions for the evaluation of seismi-
ytical solutions. Indeed, the use of an elastoplastic model is not cally induced loading and deflection in circular lined tunnels
compatible with the assumptions of the analytical solutions, while under plane strain quasi-static conditions. This was done for a
the nonlinear modelling of the excavation process results to a wide range of flexibility ratios and a variation of slippage condi-
stress-field around the tunnel which differs from the assumption tions. Based on this parametric investigation the following con-
of the analytical solutions of a uniform strain field around the clusions can be drawn:
tunnel. This was further investigated herein by repeating the FE
analysis simulating nonlinearly the excavation process and the  Assuming full-slip conditions between the soil and the lining,
tunnel construction (in all previous nonlinear quasi-static analyses the results of the analytical methods, in terms of thrust loads
the excavation/construction was simulated linearly). Fig. 15 shows and bending moments, are in perfect agreement with the
the thrust and bending moment variations around the tunnel numerical predictions. The numerical results also compared
lining for PI¼ 0, PGA ¼0.2 g and Gmax ¼448.0 MPa predicted by very well with the Park et al. [18] solution when the slippage
numerical analysis both for the case of linear and nonlinear effect between the lining and the soil was considered.
excavation/construction and the corresponding no-slip Bobet [5]  In agreement with previous studies [10,15,22], the Penzien [19]
prediction. It should be clarified that the loads generated during solution for the no-slip condition underestimates the thrust
excavation/construction have been subtracted to facilitate the values and thus its use should be avoided. The thrust loads
comparison with the analytical solution. Clearly the nonlinear predicted by the extended Hoeg [11], the Park et al. [18] and
simulation for construction/excavation leads to a non-symmetric Bobet [5] approaches are in perfect agreement with the num-
distribution of the loads around the tunnel lining, while the pre- erical solution.
dicted response diverges considerably from the analytical solution  The bending moments resulting from the Park et al. [18] and
in [5]. For both numerical analyses, the maximum response is Bobet [5] methods for the no-slip case are also in very good
predicted at θ ¼ 45o to the horizontal axis. agreement with the numerical results. The bending moments
S. Kontoe et al. / Soil Dynamics and Earthquake Engineering 66 (2014) 206–219 217

Fig. 15. Variation of seismic thrust (a) and seismic bending moment around the tunnel lining (PI ¼ 0, PGA ¼ 0.2 g, Gmax ¼448.0 MPa).

predicted by Hoeg's method start diverging from the numerical The conclusions drawn from the comparison between analy-
solution even for very low flexibility ratios (i.e. for Fo1), while tical solutions and numerical analysis is similar to the linear case.
for the Park et al. [18] and Bobet [5] methods this divergence is Park et al. [18] and Bobet [5] predictions, in terms of thrust and
minor and more pronounced only for higher flexibility ratios bending moments, compare well with the numerical analysis
(F4 21.7, reaching 15.3% for F¼185.7). Generally, the deviations results. Overall the comparison for the thrust values deteriorates
between analytical and numerical solutions in terms of bend- with increasing nonlinearity (i.e. higher PGA level and lower
ing moments are increasing with increasing flexibility ratio flexibility ratios), but it is always within acceptable levels. With
showing a trend of stabilisation for larger flexibility ratios. regards to normalised tunnel deflection, the agreement between
 In terms of normalised lining deflection, the discrepancy the analytical solutions and the numerical approach is also
between the numerical predictions and the analytical solutions satisfactory with the Hoeg and Penzien solutions showing slightly
is significant for flexibility ratios less or equal to one, while the better comparison with the numerically obtained deflections.
comparison improves for higher values of flexibility ratio. For The findings of this study confirm that analytical solutions with
the full-slip case Hoeg's, Penzien's and Park et al.'s solutions equivalent linear shear stiffness and the corresponding compatible
give identical results, while Bobet's solution gives lower values shear-strain at the tunnel axis level as input parameters can be
for the entire range of flexibility ratios. For the no-slip case and successfully used for the preliminary seismic design of circular
for large flexibility ratios the solutions in [5,18] are in better tunnels. The proposed approach can lead to more accurate esti-
agreement with the numerical predictions than the remaining mation of the lining loads as soil nonlinearity is approximated in a
two solutions. consistent manner. Analytical solutions are based, of course, on
 The solution in [5] for undrained conditions was found to be in numerous assumptions and simplifications and therefore detailed
perfect agreement with the numerical predictions for the full- design should involve a more rigorous approach. For the detailed
slip case. For the no-slip case the numerical analysis predicts design, quasi-static deformation based FE nonlinear analysis can
larger values of thrust and bending moment than in [5] for low be employed as this approach is capable of dealing with some of
flexibility ratios, while this trend is reversed for F4 21.7. the limitations of the analytical solutions (e.g. realistic modelling
of the excavation and tunnel construction). Ideally though, time-
domain finite element analysis should be used employing an
The second part of this paper investigated the use of equivalent appropriate constitutive model which can capture fundamental
linear properties in analytical solutions as an approximate way of facets of soil behaviour under seismic loading.
simulating nonlinearity which was previously suggested by Wang
[25] and later employed by Hashash et al. [9] and Amorosi and
Boldini [1]. Equivalent linear site response analyses were under- Appendix
taken varying the intensity of the ground motion and the soil type
covering a wide range of shear strains. The resulting equivalent This Appendix summarises the equations employed by the four
stiffness (Geq) and effective shear strain at the level of the tunnel analytical methods for the calculation of axial force (thrust),
axis were used as input parameters for the analytical results. bending moment and normalised lining deflection for the full
The analytical solutions were validated for the no-slip case against and no-slip assumption at the interface. The equations are pre-
FE quasi-static nonlinear analyses subjected to the same value of sented using a consistent notation for all methods, where Em and
shear-strain, but starting the loading with the elastic properties El are Young's moduli of the medium and the lining respectively,
(Gmax). νm and νl are Poisson's ratios of the medium and the lining
218 S. Kontoe et al. / Soil Dynamics and Earthquake Engineering 66 (2014) 206–219

 υm Þ 3El Ið3  4υm Þ


respectively, r and t are the radius and the thickness of the lining, where R ¼ 7 4ð1
ðα þ 1Þ and α ¼ r 3 G ð1  ν2 Þ
n
m l
respectively, θ is the angular location of the tunnel lining, I is the Δdl
moment of inertia per unit width of the lining, γmax is the ¼R ð14Þ
rγ max
maximum free-field shear-strain at the level of the tunnels axis
and C and F are the compressibility and flexibility ratios defined by
Eqs. (1) and (2) respectively.

Park et al. method [18]


Generalised Hoeg's method (often referred to Wang [25])
Park et al. [18] introduced the shear flexibility coefficient, D in
order to improve the modelling of the slip between the lining and
the soil. The equations that they propose for the calculation of the
seismic induced thrusts and bending moments are given in
(a) Full-slip
relation to this shear coefficient, with D ¼0 and D-1 represent-
1 Em  π
M ¼  K1 r 2 γ max cos 2 θ þ ð5Þ ing the no-slip and full-slip cases respectively:
6 ð1 þ νm Þ 4   
ð1  νm ÞEm γ max r 4Em π
 T¼ 2F þ ð1  2νm ÞC þ 4D cos 2 θ þ
1 Em π ð1 þ νm ÞΔ '' rð1 þ νm Þ 4
T ¼  K1 rγ max cos 2 θ þ ð6Þ
6 ð1 þ νm Þ 4 ð15Þ
 νm Þ
where K 1 ¼ ð2F12ð1
þ 5  6νm Þ   
ð1  νm ÞEm γ max r 2 4Em π
Δdl 1 M¼  ð1  2νm ÞC þ2 þ D cos 2 θ þ
¼ 7 K 1 Fγ max ð7Þ rð1 þ νm Þ
Δdf reef ield 3 ð1 þν ÞΔ''
m
4
ð16Þ

(b) No-slip where


Em π
T ¼ K 2 rγ cos 2ðθ þ Þ ð8Þ Δ'' ¼ CFð1  2νm Þ þFð3  2νm Þ
2ð1 þ νm Þ max 4
ð2F þ 5  6νm ÞEm
þCð2:5  8νm þ 6ν2m Þ þ 6  8νm þ 2D
rð1 þ νm Þ
 
Δdl 4ð1  νm Þ 8Gs FD
where ¼ ð1  2νm ÞFC þ 2F þ ð17Þ
Δdf reef ield Δ'' r
Fð1  2νm Þð1  CÞ  21 ð1  2νm Þ2 þ 2
K2 ¼ 1þ 
F½ð3  2νm Þ þ ð1  2νm ÞC þ C 52  8νm þ 6νm 2 þ 6  8νm
Bobet method [5]
It should be noted that no solutions are proposed for the calcula-
tion of the bending moment and the normalised lining deflection
for the no-slip assumption, but the method suggests using the
corresponding equations for the full-slip assumption. (a) Full-slip – dry soil
12ð1  νm Þ
T¼ Gm γ max r sin 2θ ð18Þ
3ð5  6νm Þ þ ð1  νm ÞF '
Penzien [19] method
M¼T r ð19Þ
with F ' ¼ Em r 3 ð1 ν2m Þ=El Ið1  ν2l Þ

(a) Full-slip Δdl 2U r


¼ ð20Þ
 Δdf reef ield rγ max
3E IRn r γ max π
T ¼  l3 cos 2 θ þ ð9Þ
2 r ð1  νl Þ
2 4 where U r is the radial displacement of the lining, which ignor-
ing the thickness of the lining is defined as
3El IRn r γ max  π
M¼  cos 2 θ þ ð10Þ 1 þ νm
2 r 2 ð1  ν2l Þ 4 Ur ¼ ½1 þ 4ð1  νm ÞC 1 þ C 2 Gm γ max r sin 2θ ð21Þ
Em
3  ð1  νm ÞF ' '
where with C 1 ¼  , C 2 ¼  3ð1  2νm ÞF þ ð1  νm Þ
3ð5  6νm Þ þ ð1  νm ÞF ' 3ð5  6νm Þ þ ð1  νm ÞF '
αn ¼ 3 El I ð5  6 υm Þ=2 r 3 Gm ð1  ν2l Þ, Rn ¼ 7 4ðαð1n  υm Þ
þ 1Þ , and Gm is
the shear modulus of the medium. (b) No-slip – dry soil
Δdl T ¼ ð1 C 2 ÞGm γ max r sin 2θ ð22Þ
¼ Rn ð11Þ
r γ max
1
M ¼  ð1 þC 1 þ C 2 ÞGm γ max r 2 sin 2θ ð23Þ
2
(b) No-slip ð1  νm Þ C ' þ ð1  νm Þ  ½ð1  νm ÞC ' þ 43=F '
2
where C 1 ¼  2
 ð1  ν2 ÞC ' þ ð1  νm Þð3  2νm Þ þ ½ð1  νm Þð5  6νm ÞC ' þ 4ð3  4νm Þ3=F '
3El I R Δdf reef ield π m
T¼
r ð1  νl Þ
3 2
cos 2 θ þ
4
ð12Þ 1 ð1  νm ÞC ' 2  C 1 ½ð1  νm ÞC ' þ 4νm 
C2 ¼
3 ð1  ν ÞC ' þ 2
m
3El I RΔdf reef ield  π
M¼  cos 2 θ þ ð13Þ with C ' ¼ Em rð1 ν2m Þ=El Að1  ν2l Þ and F ' ¼ Em r 3 ð1  ν2m Þ=El I
2 r ð1  νl Þ
2 2 4 ð1  ν2l Þ
S. Kontoe et al. / Soil Dynamics and Earthquake Engineering 66 (2014) 206–219 219

and [4] Bobet A. Effect of pore water pressure on tunnel support during static and
seismic loading. Tunn Undergr Space Technol 2003;18:377–93.
1 þ νm [5] Bobet A. Drained and undrained response of deep tunnels subjected to far-
Ur ¼ ½1  2ð1  νm ÞC 1  C 2 Gm γ max r sin 2θ ð24Þ
Em field shear loading. Tunn Undergr Space Technol 2010;25:21–31.
[6] Corigliano M, Scandella L, Lai CG, Paolucci R. Seismic analysis of deep tunnels
in near field fault conditions: a case study in Southern Italy. Bull Earthq Eng
(c) Full-slip – undrained conditions 2011;9(4):975–95.
[7] Darendeli MB. Development of a new family of normalised modulus reduction
6
T¼ Gm γ max r sin 2θ ð25Þ and material damping curves. ([Ph.D. thesis]). Austin: University of Texas;
6 þ ð1  νm ÞF ' 2001.
[8] Day RA, Potts DM. Zero thickness interface elements-numerical stability and
application. Int J Numer Anal Methods Geomech 1994;18:689–708.
M¼T r ð26Þ
[9] Hashash YMA, Hook JJ, Schmidt B, Yao JI-C. Seismic design and analysis of
underground structures. Tunn Undergr Space Technol 2001;16:247–93.
1 þ νm [10] Hashash YMA, Park D, Yao JI-C. Ovaling deformations of circular tunnels under
Ur ¼ ½1 þ 2C 1 þ C 2 Gm γ max r sin 2θ ð27Þ
Em seismic loading, and update on seismic design and analysis of underground
structures. Tunn Undergr Space Technol 2005;20:435–41.
with [11] Hoeg K. Stresses against underground structural cylinders. J Soil Mech Found
' '
C 1 ¼  3  ð1  νm ÞF and C 2 ¼  ð1  νm ÞF Div ASCE 1968;94(4):833–58.
6 þ ð1  νm ÞF ' 6 þ ð1  νm ÞF ' [12] Idriss IM, Sun JI. User's manual for SHAKE91. Davis: Center for Geotechnical
Modeling, Department of Civil Engineering, University of California; 1992.
(d) No-slip – undrained conditions [13] Kontoe S, Pelecanos L, Potts DM. An important pitfall of pseudo-static finite
T ¼  ð1  2C 2 ÞGm γ max r sin 2θ ð28Þ element analysis. Comput Geotech 2013;48:41–50.
[14] Kontoe S, Zdravković L, Potts DM, Menkiti CO. On the relative merits of simple
and advanced constitutive models in dynamic analysis of tunnels. Geotechni-
1
M ¼  ð1 þ 2C 1 þ 2C 2 ÞGm γ max r 2 sin 2θ ð29Þ que0016-8505 2011;61:815–29.
2 [15] Kontoe S, Zdravković L, Potts DM, Menkiti CO. Case study on seismic tunnel
with response. Can Geotech J 2008;45(12):1743–64.
[16] Kouretzis GP, Sloan SW, Carter JP. Effect of interface friction on tunnel liner
ð1  νm Þ2 C ' þ ð1  νm Þ  ½ð1 νm ÞC ' þ 43=F ' internal forces due to seismic S- and P- wave propagation. Soil Dyn Earthq Eng
C1 ¼  2013;46:41–51.
½2 þ ð1  ν ÞC ' ½ð1 ν Þ þ6=F ' 
m m [17] Owen GN, Scholl RE. Earthquake engineering of large underground structures.
(Report no. FHWA/RD-80/195). Washington, D.C: Federal Highway Adminis-
tration and National Science Foundation; 1981.
1 ð1  νm Þ2 C '  12=F ' [18] Park K-H, Tantayopin K, Tontavanich B, Owatsiriwong A. Analytical solution for
C2 ¼
2 ½2 þ ð1  ν ÞC ' ½ð1 ν Þ þ 6=F '  seismic-induced ovaling of circular tunnel lining under no-slip interface
m m
conditions: a revisit. Tunn Undergr Space Technol 2009;24:231–5.
and [19] Penzien J. Seismically induced racking of tunnel linings. Int J Earthq Eng Struct
Dyn 2000;29:683–91.
1 þ νm [20] Potts DM, Zdravković LT. Finite element analysis in geotechnical engineering:
Ur ¼ ½1  2C 1  2C 2 Gm γ max r sin 2θ ð30Þ theory. London: Thomas Telford; 1999.
Em
[21] Potts DM, Zdravković LT. Finite element analysis in geotechnical engineering:
application. London: Thomas Telford; 2001.
[22] Sedarat H, Kozak A, Hashash YMA, Shamsabadi A, Krimotat A. Contact
interface in seismic analysis of circular tunnels. Tunn Undergr Space Technol
2009;24:482–90.
References [23] Taborda D. Development of constitutive models for application in soil
dynamics. ([Ph.D. thesis]). London: Imperial College; 2011.
[24] Taborda D, Kontoe S, Zdravković L & Potts DM 2009. Application of cyclic
[1] Amorosi A, Boldini D. Numerical modeling of the transverse dynamic behavior
nonlinear elastic models to site response analysis. In: Proceedings of the 1st
of circular tunnels in clayey soils. Soil Dyn Earthq Eng 2009;29:1059–72.
international symposium on computational geomechanics, Juan-les-Pins,
[2] Argyroudis SA, Pitilakis KD. Seismic fragility curves of shallow tunnels in
alluvial deposits. Soil Dyn Earthq Eng 2012;35:1–12. France; 29 April–1 May 2009. p. 956–66.
[3] Bardet JP, Ichii K, Lin CH. EERA, a computer program for equivalent linear [25] Wang JN. Seismic design of tunnels: a state-of-the-art approach. (Monograph).
earthquake site response analysis of layered soils deposits. Los Angeles: New York: Parsons, Brinckerhoff, Quade and Douglas Inc.; 1993.
University of Southern California; 2000.

Vous aimerez peut-être aussi