You are on page 1of 27

This article was downloaded by: [National Sun Yat-Sen University]

On: 04 January 2015, At: 11:30


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

European Journal of Environmental and


Civil Engineering
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/tece20

Experimental and numerical study on


circular tunnels under seismic loading
a b b
Giovanni Lanzano , Emilio Bilotta , Gianpiero Russo & Francesco
b
Silvestri
a
DiBiT, University of Molise, Campobasso, Italy
b
DICEA, University of Napoli Federico II, Naples, Italy
Published online: 06 Mar 2014.

Click for updates

To cite this article: Giovanni Lanzano, Emilio Bilotta, Gianpiero Russo & Francesco Silvestri (2014):
Experimental and numerical study on circular tunnels under seismic loading, European Journal of
Environmental and Civil Engineering, DOI: 10.1080/19648189.2014.893211

To link to this article: http://dx.doi.org/10.1080/19648189.2014.893211

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015
European Journal of Environmental and Civil Engineering, 2014
http://dx.doi.org/10.1080/19648189.2014.893211

Experimental and numerical study on circular tunnels under seismic


loading
Giovanni Lanzanoa, Emilio Bilottab*, Gianpiero Russob and Francesco Silvestrib
a
DiBiT, University of Molise, Campobasso, Italy; bDICEA, University of Napoli Federico II,
Naples, Italy
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

(Received 9 May 2013; accepted 21 January 2014)

This paper compares the experimental results of a set of centrifuge models of tunnels
in sand under seismic loadings with the predictions of finite element dynamic analy-
ses and of simplified methods. In order to characterise the soil behaviour, mobilised
shear stiffness and damping ratio of the sand model have been back-calculated from
the experimental results according to two different procedures. Starting from the
accelerometer measurements, one was based on the transfer functions from surface to
base and the other one on the average shear stress–strain cycles along the sand layer.
A series of viscoelastic 2D dynamic analyses were performed to simulate the model
tests by a linear equivalent approach. The equivalent shear stiffness and damping
ratio determined from stress–strain cycles were used as input values for the analyses.
The shear stress transfer at the ground-lining interface was back-analysed to calibrate
the interface elements used in the numerical code, in order to improve the assessment
of the transient changes of hoop force. Finally, the numerical results have been com-
pared to analytical solutions, widely adopted in the design, and to the experimental
data in terms of transient increments of internal forces in the lining. Such a compari-
son indicates that the analytical formulations give a good estimation of the seismic
increment of bending moment in the lining and a reasonable lower bound for the
transient changes of hoop forces, provided that cyclic shear strains are correctly eval-
uated.
Keywords: centrifuge; tunnel; numerical analysis; back-analysis

1. Introduction
Shallow circular tunnels in soft ground are largely used for every kind of transportation
systems. The seismic response of these structures is generally safer compared to above
ground structures; nevertheless, several tunnels suffered strong damage in recent earth-
quakes (Bäckblom & Munier, 2002; Wang et al., 2001; Yoshida, 2009), which may be
associated with the onset of loading conditions incompatible with the lining resistance.
The increments of internal forces induced in a tunnel lining during earthquakes can
be predicted with several procedures at different levels of complexity (e.g. Bilotta et al.,
2007; Hashash, Hook, Schmidt, & Yao, 2001). For preliminary and intermediate design
stages, pseudo-static and simplified dynamic analyses are suggested by guidelines (AGI,
2005; ISO 3010:2001, 2001). These methods permit to calculate the transient changes
of internal forces in the lining once the seismic increment of shear strain at the tunnel
depth is estimated (Penzien & Wu, 1998; Wang, 1993).

*Corresponding author. Email: bilotta@unina.it

© 2014 Taylor & Francis


2 G. Lanzano et al.

On the other hand, experimental and numerical evidences of permanent changes of


internal loads in the tunnel lining (e.g. Amorosi & Boldini, 2009; Lanzano, Bilotta,
Russo, Silvestri, & Madabhushi, 2012; O’Rourke, Goh, Menkiti, & Mair, 2001) suggest
that soil-structure interaction (SSI) should be ideally modelled through full dynamic
analyses accounting for soil plastic straining under cyclic loads to capture such effects.
However, these approaches require advanced soil characterisation and numerical tools
which are not conventionally used in the design practice.
The calibration of all the above methods should require validation against experi-
mental data, which are seldom available since measurements of seismic internal forces
on real-scale structures during earthquakes are very difficult. Complication arises not
only due to the random occurrence of earthquakes, but also because the routine monitor-
ing instrumentation of the existing tunnels has too large time sampling. Hence, it is not
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

able to catch the transient nature of dynamic soil-tunnel kinematic interaction.


Due to the substantial lack of well-documented full-scale case histories, any of the pro-
posed design methods is difficult to be validated. To bridge this gap, centrifuge seismic
tests on a model tunnel (Lanzano et al., 2012) were carried out at the Schofield Centre of
Cambridge University (UK) in the framework of an Italian research project (ReLUIS).
The final aim of that work was to calibrate and compare numerical predictions of different
complexity on the basis of reliable experimental data (Bilotta & Silvestri, 2012).
A primary scope of the study presented in this paper was to interpret the experimen-
tal data from the centrifuge tests (Section 2), in order to complement the mechanical
characterisation of the sand from the laboratory with the soil properties measured at the
model scale (Section 3).
The second part of the paper (Section 4) compares the experimental results with the
predictions of simplified approaches (i.e. those suggested by Wang, 1993) and SSI
analyses. These latter were targeted to capture only the transient behaviour during shaking,
poorly affected by other well-known mechanical phenomena, such as sand densification
during shaking, which is expected to primarily induce permanent changes of internal
forces. A simple homogeneous viscoelastic model, calibrated on the basis of equivalent
linear soil parameters back-figured from the centrifuge tests, was therefore adopted.

2. Reference centrifuge tests


In this section, a brief summary of the centrifuge tests on tunnel models is given. More
explanatory details about the equipment, measuring instruments, materials, test prepara-
tion, procedures and results are reported by Lanzano et al. (2012).
The models were made by using dry Leighton Buzzard sand (fraction E); the sandy
layers were deposited at two different relative densities (Dr = 40 and 75%).
The static and dynamic properties of the sand were already measured during a com-
prehensive laboratory investigation performed on specimens prepared at comparable
densities (Visone & Santucci de Magistris, 2009). Nevertheless, a back-calculation of
the actual dynamic properties in the centrifuge model has been preferred in this study,
as discussed in Section 3.
The tunnel lining was physically modelled by an aluminium–copper alloy tube
(density, ρ = 2770 kg/m3; Young’s modulus, E = 70 GPa; Poisson’s ratio ν = .33) having
an external diameter D = 75 mm and a thickness t = .5 mm (Figure 1).
The basic features of the four series of centrifuge tests are summarised in Table 1.
The tube was located at two different depths; the layouts of the tests are shown in
Figure 2. Further details can be found in Lanzano et al. (2012).
European Journal of Environmental and Civil Engineering 3

Figure 1. Aluminium alloy tube and drawing of strain gauges layout.


Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

Table 1. Tested models.


Model Tunnel cover, C [mm] Relative density, Dr [%]
T1 75 ~75
T2 75 ~40
T3 150 ~75
T4 150 ~40

Figure 2. Instrumentation layout of models (Lanzano et al., 2012).

The tests were performed after accelerating the models at 80 g. According to the
centrifuge scale factors, each model tunnel reproduces the behaviour of a prototype con-
crete tunnel with a diameter of 6 m and a thickness of .06 m embedded in a soil layer
about 24 m thick.
Accelerometers were installed along three vertical arrays (i.e. “tunnel”, “free-field”
and “reference” in Figure 2) to measure the horizontal and the vertical components of
4 G. Lanzano et al.

the acceleration at the base of the model, on the container and in the ground. One accel-
erometer of the reference array was located on the base plate, in order to record the
applied motion at the rigid basement.
Strain gauges along the tube were arranged in order to measure bending moments
and hoop stresses at four locations along two transverse sections (see Figure 1).
The input motions were pseudo-harmonic and they had similar features in the four
models. Table 2 shows the values of amplitude, nominal frequency and duration of each
signal both at model and prototype scale (bracketed figures). As an example, the time
histories of the acceleration applied to the model T3 and the corresponding Fourier
spectra are shown in Figure 3.
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

3. Evaluation of soil properties


The fabric of sand in the centrifuge models might have been different from the labora-
tory specimens, since the procedure of preparation of the centrifuge models could not
reproduce exactly that adopted for the laboratory element tests. On the other hand, the
laboratory stress paths could not reproduce all the possible loading paths experienced by
the model during shaking. Therefore, the dynamic properties of the sand to be used in
the numerical analyses were back-figured from the results of the centrifuge tests as
shown in the following.
The reference shear strain, γ, the mobilised shear stiffness, G, and the damping ratio,
D, of the soil during each shaking were estimated following two different approaches
starting from the acceleration time histories. The two procedures are hereafter described
and the results are discussed and compared.

3.1. Coherence and transfer functions


The “similarity” between two time histories of horizontal accelerations measured in dif-
ferent points ( j, k) of the same array can be represented by the coherence function,
Cohjk(ω):
Sjk ðxÞ
Cohjk ðxÞ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (1)
Sjj ðxÞSkk ðxÞ
where Sjk is the Fourier transform of the cross-covariance and Sjj and Skk are the auto-
spectra of the Fourier transform of the auto-covariance. For each frequency, the coher-
ence has a value included between 0 and 1: the higher the coherence, the higher the cor-
relation between the two signals.
The transfer function, instead, is a representation of the ground motion variability
due to the wave propagation inside a medium. This function is defined as the ratio

Table 2. Model earthquakes.


Frequency
Input signal Gravity level [g] [Hz] Duration [s] Amplitude [g]
EQ1 80 30 (.375) .4 (32) 4.0 (.05)
EQ2 80 40 (.5) .4 (32) 8.0 (.10)
EQ3 80 50 (.625) .4 (32) 9.6 (.12)
EQ4 80 60 (.75) .4 (32) 12.0 (.15)
European Journal of Environmental and Civil Engineering 5
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

Figure 3. Input signal and smoothed Fourier spectra at model scale recorded on the base-plate
(T3; 80 g).

between the Fourier spectra, X(ω) and Y(ω), of two different signals. The modulus of
the transfer function represents the amplification function, A(ω), which is defined as:
X ðxÞ
AðxÞ ¼ jHðxÞj ¼ (2)
Y ðxÞ
The combined interpretation of the transfer and the coherency functions was used in this
case to recognise which frequencies were amplified in the waves propagating through a
soil layer between two accelerometers. Figure 4 shows all the transfer and coherence
functions which were calculated along the vertical reference arrays located in the four
models, between the base, Y(ω), and the top accelerometers, X(ω).
In the range of frequencies where the coherence was higher than .9, it was assumed
that the intensity of the recorded signal was about one order of magnitude higher than
any possible noise. Hence, experimental values in this range were best fitted with the
theoretical amplification function of a viscoelastic layer, i.e:
1
AðxÞ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (3)
cos2 F þ ðDFÞ2

where D is the soil damping ratio and F is the frequency ratio, defined as
F ¼ x  H=VS , being H the thickness of the sand layer and VS the shear wave velocity.
In most cases, the two signals, X(ω) and Y(ω), met the requirement on the minimum
coherence value through more than 50% of the frequency range, and the coefficient of
determination of the best fitting, R2, was higher than .9.
Similar curves were fitted also on the basis of experimental data along free-field and
tunnel verticals (Bilotta, Lanzano, Russo, Silvestri, & Madabhushi, 2009; Lanzano,
Bilotta, Russo, Silvestri, & Madabhushi, 2009).
6 G. Lanzano et al.
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

Figure 4. Coherence and transfer functions relative to reference vertical.

In this way, the fundamental frequency of the layer:


f1 ¼ x1 =2p ¼ VS =4H (4)
and the mobilised damping ratio, D, was obtained from the peaks of the back-figured
amplification functions. The mobilised transversal modulus of elasticity, G, was evalu-
ated from f1 using the expression:
GTF ¼ VS2 q ¼ ð4Hf1 Þ2 q (5)
being ρ the soil mass density.
An average value of the shear strain mobilised during each earthquake can be calcu-
lated according to Newmark (1967):
vmax
cTF ¼ (6)
VS
In Equations (5) and (6), VS is the equivalent shear wave velocity obtained from the
transfer function through Equation (4), and vmax is the average peak particle velocity in
the soil layer. The latter was back-figured from the integration of the acceleration time
history recorded by the central accelerometer of the reference array (Conti &
Viggiani, 2012).
European Journal of Environmental and Civil Engineering 7

The mobilised values of the dynamic shear modulus, the damping ratio and the
maximum shear strain are reported in Table 3. The average values of the dynamic shear
modulus range between 28 and 10 MPa, from EQ1 to EQ4, confirming that the mobi-
lised stiffness of the soil was relatively low. Some values in the table are reported in
brackets, since in these cases, either the best fitting was only possible on a partial span
of the data points (lower than 25%) or the determination coefficient R2 was lower than
.85. These values are considered as not reliable.
The average back-figured values of the damping ratio are rather high (14–35%);
however, they could well be overestimated due to low coherence of the signals in the
frequency range around the resonance peak.
The average values of the maximum shear strain range between .07 and .23%,
without any clear difference between dense and loose sand.
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

3.2. Shear stress–strain loops


Zeghal and Elgamal (1994) suggested a procedure to evaluate the shear modulus, the
damping ratio and the shear strain from the data recorded by arrays of accelerometers.
Brennan, Thusyanthan, and Madabhushi (2005) adapted the method to back-figure the
same parameters in dynamic centrifuge tests; the procedure has been more recently vali-
dated also by Li, Escoffier, and Kotronis (2013). Following the same approach, the time
histories of the displacements, u(t), were obtained from double integration of the accel-
erograms, a(t). In order to avoid errors like an unrealistic drift during the shaking, the
signal was band-pass filtered twice (between 15 and 250 Hz), before each integration.
The filter also eliminated the phase distortion due to the integration procedure.
The time histories of shear strain could be calculated by differentiating the displace-
ments, u(t), with respect to depth, z, using a second-order approximation between two
or more instruments positioned along the same vertical array:

Table 3. Mobilised shear stiffness, damping ratio and maximum shear strain from transfer
functions.
Model EQ1 EQ2 EQ3 EQ4
GTF [MPa]
T1 31 37 39 34
T2 27 45 (21) (10)
T3 32 29 (25) (10)
T4 23 23 (24) (11)
Avg 28 33 27 16
DTF [%]
T1 14 20 28 21
T2 8 14 (28) (40)
T3 17 19 (16) (39)
T4 17 14 (14) (40)
Avg 14 17 22 35
γTF [%]
T1 .07 .08 .11 .15
T2 .07 .07 (.15) (.26)
T3 .07 .08 (.13) (.27)
T4 .08 .09 (.14) (.25)
Avg .07 .08 .13 .23
8 G. Lanzano et al.
h i
ðzi zi1 Þ ðziþ1 zi Þ
ðuiþ1  ui Þ ðziþ1 zi Þ
þ ðui  ui1 Þ ðz i zi1 Þ
cðzi Þ ¼ (7)
ðziþ1  zi1 Þ
in which the index i was relative to the position of the central instrument and i – 1 and
i + 1 to the upper and lower accelerometer, respectively.
The shear stress τ was computed through the dynamic equilibrium of a soil column,
by integrating the acceleration time histories with respect to depth:
Z z
sðzÞ ¼ qadz (8)
0

As the signals were not of single frequency, the shape of the cycles was not regular (cf.
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

also Brennan et al., 2005; Conti & Viggiani, 2012; Li et al., 2013). An equivalent shear
stiffness, Gcyc, for each cycle was generally estimated as:
smax  smin
Gcyc ¼ (9)
cmax  cmin
The equivalent damping ratio, Dcyc, was calculated by integrating the area of each
stress–strain loop as follows:
H
2 sdc
Dcyc ¼ (10)
p Ds  Dc
where Δτ and Δγ are the peak-to-peak amplitudes of the shear stress and strain in each
loop. Table 4 shows the values of Gcyc and Dcyc averaged over the stress–strain cycles.
The single amplitudes of cyclic shear strain were rather high, varying between .02
and .3%, which correspond to a strain range in which the stress–strain behaviour of
sand is clearly non-linear. As a consequence, the equivalent shear stiffness decreases
from values as high as 41 MPa (T1, EQ1) to 11 MPa (T2, EQ4); accordingly, the
equivalent damping increases, ranging between 11 and 38%.

3.3. Comparison between the two procedures


Figure 5 compares the peak shear strains and the equivalent parameters calculated
according to the two procedures above described. The subscript “TF” refers to the
values computed using the transfer functions for G and D, and Equation (6) for the
shear strain γ; the subscript “cyc” is used for the values obtained from the τ – γ cycles.
The graphs show a reasonable good agreement between the two methods. The val-
ues computed for dense (T1, T3) and loose (T2, T4) soil are plotted with full and open
symbols, respectively. The interpretation based on the transfer functions produces higher
values for both the shear strain, γTF, and stiffness, GTF. It may be also observed that the
interpretation of the cycles shows slightly larger strains, γcyc, for the looser models than
for the denser ones (Figure 5(a)). Correspondingly, the shear stiffness values, Gcyc, of
the looser models are lower (Figure 5(b)).
It is worth noting that the quality of the results based on the transfer function is
affected by the use of an almost single-frequency shaking event. In fact, in most cases,
the frequency of the input signal was far from the fundamental frequency of the soil layer,
hence the most amplified harmonics of the signal were associated to a very low energy
content. The use of a shaking signal with a wider range of frequencies, as natural
earthquakes are, would likely improve the reliability of the transfer function method.
European Journal of Environmental and Civil Engineering 9

Table 4. Mobilised shear stiffness, damping ratio and maximum shear strain from stress–strain
cycles.
Model EQ1 EQ2 EQ3 EQ4
Gcyc [MPa]
T1 41 30 26 21
T2 22 22 14 11
T3 21 20 20 14
T4 22 16 14 13
Avg 30 22 19 15
Dcyc [%]
T1 17 20 20 28
T2 14 27 32 38
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

T3 12 19 22 28
T4 16 11 28 31
Avg 14 19 25 31
γcyc [%]
T1 .02 .05 .09 .13
T2 .04 .07 .15 .26
T3 .04 .06 .09 .18
T4 .04 .09 .13 .21
Avg .03 .07 .11 .19

Figure 5. Comparison between the two different estimations of (a) γ, (b) G and (c) D.

The comparison in Figure 5(c) between the two methods in terms of the damping
ratio, D, is affected by the larger scatter in the results, as also observed in similar works
(cf. also Brennan et al., 2005; Conti & Viggiani, 2012; Li et al., 2013).

3.4. Strain dependence of shear stiffness and damping


The small strain shear modulus, G0, of the LB sand was measured at varying values of
confining pressure and relative density in the low-amplitude resonant column (RC) and
torsional shear (TS) tests on laboratory specimens (Visone & Santucci de Magistris,
2009). However, for the reasons mentioned at the beginning of this section, a direct
assessment of G0 from the back-analysis of the centrifuge tests was preferred in this
work.
Brennan et al. (2005) suggested to evaluate the initial soil stiffness of the centrifuge
model by a low-energy dynamic shaking test, associated to a very low deformation
10 G. Lanzano et al.

level. More recently, Lee, Wang, Wei, and Hung (2012) have further demonstrated that
the values of shear wave velocity, VS, obtained from such low-intensity shaking tests
agree with the measurements in bender element tests during the centrifuge flight. Unfor-
tunately, even the weakest model earthquake (EQ1) induced shear strains too high to be
associated to the initial value of the shear modulus, G0. Therefore, this parameter was
obtained by extrapolation from the experimental data according to the following
procedure:

(1) Preliminarily, a mean “laboratory” G(γ)/G0 decay curve (solid line in Figure 6)
was fitted on the data of G(γ)/G0 from RC and TS tests (Lanzano, Bilotta,
Russo, Silvestri, & Madabhushi, 2010); the solid line was fitted on both RC and
TS laboratory tests data carried out at 100 and 200 kPa of confining pressure
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

(Visone & Santucci de Magistris, 2009), being such stress range representative
of the stress level in the sand model at 80 g;
(2) centrifuge experimental data points (G:γ) were obtained for both loose and dense
sand centrifuge models from the (τ, γ) cycles;
(3) reference values of shear modulus G(γmin) were determined (see Figure 7) corre-
sponding to the lowest measured shear strain γmin in centrifuge (about .02% for
the denser and .04% for the looser soil models);
(4) G(γmin) values were hence used to plot dimensionless “centrifuge” decay curves,
G/G(γmin);
(5) in order to extrapolate the value of G0 in the centrifuge model, these curves
were subsequently scaled down to match the “laboratory” G(γ)/G0 decay curve;
(6) since the necessary scaling factor is equal to G(γmin)/G0, the “centrifuge” values
of G0 could be back-figured.

On average, G0 resulted equal to 60 MPa for the denser models and to 30 MPa for
the looser ones. They result lower than the corresponding values from laboratory ele-
ment tests, which induced to infer that the overall fabric of sand in centrifuge models
was different from the laboratory specimens.
At low strain level, the calculated shear moduli were affected by large scatter, there-
fore for the lowest amplitude events (i.e. EQ1 of all models), an average value of G/G0
over all the cycles was plotted in Figure 6(a). In the case of the strongest events (e.g.
EQ4), a decay of the shear modulus was observed during the event and data from each
single cycle were plotted in the same figure.
A similar procedure was followed by using the experimental transfer functions
shown in Figure 4; the corresponding data are shown in Figure 6(b). In this case, values
of G0 equal to about 50 MPa for the densest models and 43 MPa for the loosest ones
were obtained. In the figure, the points corresponding to the events EQ4 of models T2–
T4 have been highlighted; since the procedure was applied to a very small portion of
the transfer function, the results should be considered as less reliable (cf. §3.1).
In Figure 8, a comparison between laboratory and centrifuge tests is shown in terms
of damping ratio D.
Overall, the values of damping ratio D obtained from the centrifuge (τ, γ) cycles are
relatively higher than the mean D – γ curve measured in the laboratory.
The observation of lower stiffness and higher damping in the centrifuge models con-
currently suggest that the model preparation and stress-induced anisotropy considerably
affected the fabric of the sand layer making it weaker than the laboratory specimens.
European Journal of Environmental and Civil Engineering 11
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

Figure 6. G(γ)/G0 laboratory curve compared with the values back-figured from stress–strain
cycles (a) and transfer functions (b) in centrifuge tests.

Figure 7. Procedure to compare the decay of stiffness with strain as obtained from laboratory
and centrifuge tests.
12 G. Lanzano et al.
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

Figure 8. Damping ratio laboratory curve compared with the values back-figured from
stress–strain cycles (a) and transfer functions (b) in centrifuge tests.

4. Experimental vs. numerical and analytical modelling


4.1. Equivalent linear viscoelastic analyses
Equivalent linear viscoelastic dynamic analyses of the coupled ground-tunnel system
undergoing shaking were performed by the FE code PLAXIS 2D (Brinkgreve, Swolfs,
& Engin, 2011). The geometry of the centrifuge models was reproduced by the finite
element meshes shown in Figure 9, for both shallow and deep tunnels. The PLAXIS
mesh consisted of triangular 15-node elements. The elements were small enough to
ensure the frequency content of the input signal not to be artificially filtered
(Kuhlemeyer & Lysmer, 1973). The lining was modelled by using beam elements. The
two vertical boundaries were linked by rigid “node-to-node anchors”, forcing them to
have identical displacements as in the rigid laminar box in the centrifuge tests. The
vertical spacing of the anchors is .01 m; hence, 29 constraints were used in the meshes.
A rigid bottom boundary was assumed in the FE models and the signal recorded at the
base of the reference array was assumed as input signal for the analyses.
In the numerical simulations shown hereafter, the soil was intentionally modelled as
linear viscoelastic and homogeneous, by assuming the shear stiffness G and the damp-
ing ratio D constant with depth, and corresponding to the values back-figured from the
τ – γ cycles (Table 4). Local changes of density were not modelled in the analyses. This
mimics the assumptions at the base of the simplified analytical methods such as that
proposed by Wang (1993). In PLAXIS, the viscous damping is frequency dependent
European Journal of Environmental and Civil Engineering 13
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

Figure 9. Analysed numerical models.

and traditionally modelled through the well-known Rayleigh formulation, for which the
damping matrix is evaluated as the sum of mass and stiffness matrix, multiplied for the
coefficients αR and βR, respectively. These coefficients were estimated using the “double
frequency approach” suggested by Park and Hashash (2004): the Rayleigh damping was
assumed coincident with the assigned damping ratio, Dcyc, at the predominant frequency
of the input signal (cf. Table 2) and at the first natural frequency of the soil layer. The
relevant Rayleigh functions in the frequency domain are plotted in Figure 10 for all the
analysed models, together with the Fourier spectrum of the input signals (grey shadow).
Note that the input frequency increases with the model earthquake intensity from EQ1
to EQ4 (cf. Table 2) while, correspondingly, the natural frequency decreases due to the

Figure 10. Adopted Rayleigh damping functions vs. experimental values of damping ratio.
14 G. Lanzano et al.

non-linear behaviour of sand (cf. Section 3.1); since the two frequencies approach each
other, the Rayleigh formulation produces larger overdamping at high frequencies when
passing from weaker (EQ1) to stronger (EQ4) input motions.

4.2. Dynamic response of soil layer


The time histories of accelerations were calculated at the same locations of the acceler-
ometers in the models (cf. Figure 2). In Figure 11, a comparison is shown between mea-
sured and calculated accelerations at the upper sensor along the reference array. The
curves were truncated at .23 s to allow a better insight.
The calculated time histories of acceleration for EQ3 and EQ4 are generally
smoother than the corresponding experimental ones. Compared to EQ1 and EQ2, such a
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

larger filtering of the higher frequencies is systematic in all models and it is consistent
with the numerical overdamping shown in Figure 10.
The effect of overdamping at higher frequencies can be appreciated also in Figure 12,
where a similar comparison between experimental and numerical pseudo-acceleration

Figure 11. Measured and calculated time histories of acceleration at the “reference” top acceler-
ometer (window between .03 and .23 s).
European Journal of Environmental and Civil Engineering 15

response spectra is shown. In all cases, the pseudo-acceleration is well computed around
the predominant period of the input signal. At lower periods, on the contrary, the
numerical predictions generally tend to underestimate the experimental spectral accelera-
tion. This is particularly evident for EQ3 and EQ4 in all models. The most accurate
description of the spectral response was achieved for EQ1 of model T1: in this case, the
Rayleigh function reproduced a damping value close to the measured damping ratio
over a wider range of frequencies of the input signal (cf. Figure 10).
The maximum values of measured acceleration, amax, are plotted against depth in
Figure 13, together with the corresponding calculated values.
The agreement between experimental data and numerical calculation is generally fine
along both the “reference” and the “tunnel” vertical arrays. Better predictions were
achieved for weaker shakings. This has been previously observed and discussed in terms
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

of time histories of acceleration and response spectra of the accelerometer at the top of
the reference array (Figures 11 and 12). The quality of prediction is somehow confirmed
also in depth, although the agreement with the experimental data seems higher for the
denser models (T1 and T3). A possible reason is that the experimental loose models are

Figure 12. Numerical vs. experimental response spectra of pseudo-acceleration (“reference” top
accelerometer).
16 G. Lanzano et al.
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

Figure 13. Profiles of computed and measured amax with depth.


European Journal of Environmental and Civil Engineering 17

less homogeneous than the dense ones due to the different sand deposition technique
which has been adopted in the two cases (cf. Lanzano et al., 2012).
The bottom accelerometer of the tunnel array was not connected to the box, hence it
did not record the input signal. This is the reason why the maximum values of accelera-
tion at that level differ from the corresponding ones in the reference array. Along the
tunnel array, the numerical analyses systematically underpredict these values, being clo-
ser to the maximum values of the input signals (i.e. measured at the bottom of the refer-
ence array): such a difference can be expected since the experimental detail of the
actual contact between the sand and the bottom of the box could not be modelled in the
numerical analyses.
Figure 14 illustrates the overall cyclic stress–strain behaviour of the sand model
considered as subjected to simple shear. “Experimental” and “numerical” plots are
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

Figure 14. Shear stress–strain cycles calculated from the “reference” accelerometer array (experi-
mental vs. numerical).
18 G. Lanzano et al.

compared in the figure. The “numerical” curves were obtained according to the same
procedure used for the experimental data (cf. §3.2) along the “reference” array.
On the average, the predicted shear strains result higher than experimental ones,
since the computed accelerations sometimes overestimate measurements (see for
instance response spectra in Figure 12).
The figure also shows the difficulties which may arise in back-figuring such values,
due to the irregular shape of most cycles, particularly for the weaker motions (EQ1), as
a consequence of the particular pseudo-harmonic shape of the input signal.
Moreover, the comparison between numerical and experimental cycles shows a sys-
tematic higher damping in the numerical results, as it would be expected due to the
numerical overdamping related to the Rayleigh formulation adopted.
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

4.3. Internal forces in the lining during shaking


The centrifuge tests have shown that both transient (reversible) and accumulated
(permanent) changes of internal forces arise in the lining, the latter due to sand densifi-
cation (cf. Lanzano et al., 2012). The simplified constitutive model adopted in both the
numerical and analytical predictions only allowed for the calculation of the reversible
changes of hoop forces and bending moments, due to ovalisation of the tunnel lining
under cyclic ground shear straining.

Figure 15. Non-dimensional increments of hoop forces calculated in the lining by varying the
factor Rinter.
European Journal of Environmental and Civil Engineering 19

According to the analytical formulations in literature (e.g. Penzien & Wu, 1998;
Wang, 1993), the changes of hoop forces in the tunnel lining during shaking are largely
influenced by the contact conditions at the interface between the soil and the lining.
A parametric study was therefore performed on the numerical models in order to
analyse the sensitivity of the predicted internal forces to the degree of roughness of the
soil-lining interface.

4.3.1. Back-analysis of the interface conditions


The interface elements used in the analysis were elastic, with a reduced stiffness com-
pared to the soil. In the Plaxis code, the degree of such reduction is defined by an inter-
face factor, Rinter, which is equal to 1 if no reduction is considered. In this case, no
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

relative slippage is allowed between the lining and the soil. As long as the factor Rinter
is reduced, a larger amount of relative slippage is allowed, tending to full-slip conditions
for very low values of Rinter. Four values of Rinter were used in the analyses (Rinter = .01;
.05; .1; and 1).
In Figure 15, the non-dimensional increments of hoop forces, ΔN/τR, calculated in
the lining by varying the factor Rinter are shown for all the models and all the shaking
events (τ is the reference shear stress, calculated as Gcyc · γcyc, and R is the tunnel
radius). In the same figure, the values analytically computed with the formulas by Wang
(1993) in both full-slip and no-slip conditions (see Appendix) are shown for comparison
with dashed lines. It is worth noting that, according to Wang (1993), the value of the
normalised increment of hoop force depends on the mobilised shear stiffness.

Figure 16. Measured vs. calculated changes of hoop force: influence of the choice of Rinter.
20 G. Lanzano et al.

It can be observed that all the calculations performed with Rinter = 1 predict an incre-
ment of hoop force which practically coincides with the analytical predictions for no-
slip conditions. On the other side, the calculations with Rinter = .01 are very close to the
analytical predictions for full-slip conditions.
In order to choose a reasonable value of Rinter to reproduce the fairly smooth lining–
ground interface in the centrifuge model, the experimental variations of hoop force,
ΔNexp, were compared to the corresponding numerical computations, ΔNnum, for three
different assumptions of Rinter (.01; .05; .1).
In Figure 16, the numerical changes of hoop forces, ΔNnum, are plotted against the
experimental ones, ΔNexp. The black solid line represents the condition DNnum ¼ DNexp ;
the black dashed and the grey solid lines bound a difference of ±20 and ±50%, respec-
tively, from that condition. The comparison shows that the best prediction of hoop
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

forces is obtained by assuming Rinter = .05, although for model T3 only this choice over-
estimates the experimental measurements.

Figure 17. Measured and calculated values of changes of hoop force.


European Journal of Environmental and Civil Engineering 21

4.3.2. Prediction of transient changes of internal forces


A comparison among the reversible increments of hoop forces, ΔN, and bending
moment, ΔM, is proposed in Figures 17 and 18.
In Figure 17, the experimental increments of hoop force are shown with black dots,
together with the numerical prediction for Rinter = .05 (black line) and the analytical
solution (grey line). This latter was computed by assuming in Equation (A2) (full slip,
see Appendix) γ = γcyc, i.e. the back-figured experimental values of average shear strain
in the model. The analytical solution, corresponding to full-slip conditions, is obviously
a lower bound of the experimental values. The FE analyses predict values which are in
most cases in good agreement with the experimental ones.
Figure 18 shows similar plots for bending moments, with those relevant to the
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

analytical solutions computed again with reference to the experimental strains. In this
case, the Wang’s analytical solution (Equation (A1)) does not depend on the slippage at
the interface. The differences between analytical and numerical calculations are consis-
tent with the differences between experimental and numerical shear strain previously
observed. They are however very small for the weaker events.

Figure 18. Measured and calculated values of changes of bending moment.


22 G. Lanzano et al.

5. Conclusions
In this paper, the results of centrifuge tests on reduced scale models of a circular tunnel
in sand have been interpreted and back-analysed. The aim was to compare the experi-
mental results with the predictions of simplified methods and finite element analyses.
The dynamic response of the physical models has been modelled by a simple equiv-
alent linear approach in a series of finite element analyses. This is the common proce-
dure adopted for site seismic response analysis in design, where the soil is usually
modelled as an equivalent linear viscoelastic medium. Although both experimental and
numerical evidences of permanent changes of internal loads in the tunnel lining are
reported in literature, this aspect was intentionally not examined in this work.
The mobilised shear stiffness and damping ratio were back-calculated from the
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

experimental results. The centrifuge values of small strain shear stiffness were on aver-
age lower than those measured in the laboratory element tests, due to unexpected
weaker fabric of the sand and differences in the followed loading paths.
Two procedures were used to back-figure the mobilised shear stiffness and damping
ratio from the centrifuge tests: one was based on the interpretation of the spectral ratio
between surface and bottom accelerometric signals and the other on the shear stress–
strain cycles during each event.
The degree of uncertainty of both procedures is increasingly high at the lowest strain
levels, where the input signal was most affected by high frequency noise, although time
histories were filtered for high frequencies before each integration step. The scatter of
the parameters derived from the stress–strain loops is related to the back-calculation pro-
cedure, which determines the shear strain by numerical manipulation (time integration
and spatial derivation) of the recorded accelerations.
Due to the characteristics of the input signals, which were pseudo-harmonic (i.e.
nominally characterised by a single frequency), the procedure based on the spectral ratio
allowed for results of lower reliability than the interpretation of cycles. Therefore, only
the latter was used to back-figure the mobilised stiffness and damping for the numerical
and analytical predictions.
The stiffness reduction factor of the ground-lining interface (Rinter) was back-calcu-
lated to be used in the numerical analyses in order to provide a reliable estimate of the
transient changes of hoop force. The comparison of experimental and numerical results
in terms of transient changes of internal forces in the lining was satisfactory both in
terms of bending moment and of hoop force.
The results indicate that the analytical formulations proposed by Wang (1993),
widely adopted in the design practice, can give a good estimation of the seismic incre-
ment of bending moment in the lining and a reasonable lower bound for the transient
changes of hoop forces, provided that cyclic shear strain are correctly evaluated.
Although such a simplified approach may be sufficient for a preliminary design, concern
arises when the sand layer is interested by intense densification during shaking. In this
case, permanent changes of internal forces may originate, which can be predicted only
with more sophisticated constitutive models. A careful investigation of this issue of con-
cern was however out of the scope of this work.

Acknowledgements
The experimental data at the basis of this paper were obtained in the framework of an agreement
between Universities of Napoli and Cambridge, as a part of a Research Project funded by ReLUIS
European Journal of Environmental and Civil Engineering 23

(University Network of Seismic Engineering Laboratories) Consortium. The writers wish to thank
Dr Gopal Madabhushi, who coordinated the experimental campaign in centrifuge, Dr Filippo Sant-
ucci de Magistris, for making available the characterization of LB sand by laboratory tests, and
the coordinator of the research project, Prof. Stefano Aversa, for his continuous support.

References
AGI. (2005). Guidelines for geotechnical design in seismic zones (Patron ed.). Bologna: Special
volume of Italian Geotechnical Association (in Italian).
Amorosi, A., & Boldini, D. (2009). Numerical modelling of the transverse dynamic behaviour of
circular tunnels in clayey soils. Soil Dynamics and Earthquake Engineering, 29, 1059–1072.
Bäckblom, G., & Munier, R. (2002). Effects of earthquakes on the deep repository for spent fuel
in Sweden based on case studies and preliminary model results (Technical Report TR-02-24).
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

Stockholm: Swedish Nuclear Fuel and Waste Management.


Bilotta, E., Lanzano, G., Russo, G., Santucci de Magistris, F., Aiello, V., Conte, E., … Valentino,
M. (2007). Pseudo-static and dynamic analyses of tunnels in transversal and longitudinal
direction. In Proceedings of the 4th international conference on Earthquake Geotechnical
Engineering. Thessaloniki: Springer.
Bilotta, E., Lanzano, G., Russo, G., Silvestri, F., & Madabhushi, S. P. G. (2009). Seismic analyses
of shallow tunnels by dynamic centrifuge tests and finite elements. In Proceedings of the 17th
international conference on Soil Mechanics and Geotechnical Engineering. Alexandria, Egypt:
Balkema.
Bilotta, E., & Silvestri, F. (2012). A predictive exercise on the behaviour of tunnels under seismic
actions. In Geotechnical aspects of underground construction in soft ground – Proceedings of
the 7th international symposium on Geotechnical Aspects of Underground Construction in
Soft Ground (pp. 1071–1077). Rome: Balkema.
Brennan, A. J., Thusyanthan, N. I., & Madabhushi, S. P. G. (2005). Evaluation of shear modulus
and damping in dynamic centrifuge tests. Journal of Geotechnical and Geoenvironmental
Engineering, 131, 1488–1497.
Brinkgreve, R. B. J., Swolfs, W. N., Engin, E. (2011). Plaxis 2D 2011 manuals. Delft: Plaxis bv.
Conti, R., & Viggiani, G. M. B. (2012). Evaluation of soil dynamic properties in centrifuge tests.
Journal of Geotechnical and Geoenvironmental Engineering, 138, 850–859.
Hashash, Y. M. A., Hook, J. J., Schmidt, B., & I-Chiang Yao, J. (2001). Seismic design and anal-
ysis of underground structures. Tunnelling and Underground Space Technology, 16, 247–293.
Hoeg, K. (1968). Stresses against underground structural cylinders. Journal of the Soil Mechanics
and Foundation Division, ASCE, 94, 833–858.
ISO 3010:2001. (2001). Basis for design of structures – Seismic actions on structures.
Kuhlemeyer, R. L., & Lysmer, J. (1973). Finite element method accuracy for wave propagation
problems. Journal of the Soil Mechanics and Foundation Division, ASCE, 99, 421–427.
Lanzano, G., Bilotta, E., Russo, G., Silvestri, F., & Madabhushi, S. P. G. (2010). Dynamic centri-
fuge tests on shallow tunnel models in dry sand. In Proceedings of the ICPMG2010 Physical
Modeling in Geotechnics (pp. 561–567). Zurich, Switzerland: Balkema.
Lanzano, G., Bilotta, E., Russo, G., Silvestri, F., & Madabhushi, S. P. G. (2012). Centrifuge mod-
eling of seismic loading on tunnels in sand. Geotechnical Testing Journal, 35, 854–869.
Lanzano, G., Bilotta, E., Russo, G., Silvestri, F., & Madabhushi, S. P. G. (2009). Experimental
assessment of performance-based methods for the seismic design of circular tunnels. In
Proceedings of the 1st international conference on Performance Based Design in Earthquake
Geotechnical Engineering. IS-Tokyo, Tokyo (Japan).
Lee, C.-J., Wang, C.-R., Wei, Y.-C., & Hung, W.-Y. (2012). Evolution of the shear wave velocity
during shaking modeled in centrifuge shaking table tests. Bulletin of Earthquake Engineering,
10, 401–420.
Li, Z., Escoffier, S., & Kotronis, P. (2013). Using centrifuge tests data to identify the dynamic soil
properties: Application to Fontainebleau sand. Soil Dynamics and Earthquake Engineering,
52, 77–87.
Newmark, N. M. (1967). Problems in wave propagation in soil and rocks. In Proceedings of the
international symposium on Wave Propagation and Dynamic Properties of Earth Materials
(pp. 7–26). Albuquerque, NM: University of New Mexico Press.
24 G. Lanzano et al.

O’Rourke, T. D., Goh, S. H., Menkiti, C. O., & Mair, R. J. (2001). Highway tunnel performance
during the 1999 Duzce earthquake. In Proceedings of the 15th international conference on
Soil Mechanics and Geotechnical Engineering (pp. 1365–1368). Istanbul, Turkey.
Park, D., & Hashash, Y. M. A. (2004). Soil damping formulation in nonlinear time domain site
response analysis. Journal of Earthquake Engineering, 8, 49–274.
Penzien, J., & Wu, C. L. (1998). Stresses in linings of bored tunnels. Earthquake Engineering &
Structural Dynamics, 27, 283–300.
Visone, C., & Santucci de Magistris, F. (2009). Mechanical behaviour of the Leighton Buzzard
Sand 100/170 under monotonic, cyclic and dynamic loading conditions. In Proceedings of the
XIII conference L’Ingegneria Sismica in Italia, ANIDIS. Bologna, Italy.
Wang, J.-N. (1993). Seismic design of tunnels: A state-of-the-art approach. In Monograph 7 (147 p).
New York, NY: Parsons Brinckeroff.
Wang, W. L., Wang, T. T., Su, J. J., Lin, C. H., Seng, C. R., & Huang, T. H. (2001). Assessment
of damage in mountain tunnels due to the Taiwan Chi-Chi earthquake. Tunnelling and Under-
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

ground Space Technology, 16, 133–150.


Yoshida, N. (2009). Damage to subway station during the 1995 Hyogoken-Nambu (Kobe) earth-
quake. In T. Kokusho (Ed.), Earthquake geotechnical case histories for performance-based
design (pp. 373–389). Tokyo: CRC Press.
Zeghal, M., & Elgamal, A.-W. (1994). Analysis of site liquefaction using earthquake records.
Journal of Geotechnical and Geoenvironmental Engineering ASCE, 120, 71–85.

Appendix
Wang (1993) proposed expressions to calculate the increments of bending moments and
hoop forces due to the ovalisation of a circular elastic lining (Es, νs) of thickness ts
around a cavity of radius R in an elastic ground (E, ν or G).
Assuming an average ground shear strain γ, they read as follows:
ð1 þ 3a2  4a3 Þ  p
DM ¼ cos 2 h þ  G  c  R2 (A1)
3 4

ð1 þ 3a2  4a3 Þ  p
DN ¼ cos 2 h þ GcR (full slip allowed at the interface)
3 4
(A2)
 p
DN ¼ ð1 þ a2 Þ cos 2 h þ  G  c  R (no slip allowed at the interface) (A3)
4
where θ is the angular coordinate of the lining section; a1, a2 and a3 are functions (see
Table A1) of the soil Poisson’s ratio, ν, and of two compressive and flexural relative
stiffness ratios, C and F, defined as:
Eð1  m2s ÞR
C¼ (A4)
Es ts ð1 þ mÞð1  2mÞ

Eð1  m2s ÞR3


F¼ (A5)
6Es Ið1 þ mÞ

being I the moment of inertia per unit length of lining.


European Journal of Environmental and Civil Engineering 25

Table A1. Relative stiffness parameters (Hoeg, 1968).


No slip at the interface Full slip at the interface
ð12mÞðC1Þ
a1 ¼ ð12mÞCþ1
ð12mÞð1CÞF12ð12mÞ2 Cþ2
a2 ¼ ½ð32mÞþð12mÞC Fþð528mþ6m2 ÞCþ68m
a2 ¼ 2Fþ12m
2Fþ56m
½1þð12mÞC F12ð12mÞC2
a3 ¼ ½ð32mÞþð12mÞC Fþð528mþ6m2 ÞCþ68m
a2 ¼ 2Fþ56m
2F1
Downloaded by [National Sun Yat-Sen University] at 11:30 04 January 2015

You might also like