Вы находитесь на странице: 1из 15

Al-Defae, A. H. et al. (2013). Géotechnique 63, No. 14, 1230–1244 [http://dx.doi.org/10.1680/geot.12.P.

149]

Aftershocks and the whole-life seismic performance of granular slopes


A . H . A L - D E FA E  , K . C AU C I S † a n d J. A . K NA P P E T T †

Shallow embankment slopes are commonly used to support elements of transport infrastructure in
seismic regions. In this paper, the seismic performance of such slopes in non-liquefiable granular soils
is considered, focusing on permanent movement and dynamic motion at the crest, which would form
key inputs into the aseismic design of supported infrastructure. In contrast to previous studies, the
evolution of this behaviour under multiple sequential strong ground motions is studied through
dynamic centrifuge, numerical (finite-element, FE) and analytical (sliding-block) modelling, the
centrifuge tests being used to validate the two non-physical approaches. The FE models focus on the
specification of model parameters for existing non-linear constitutive models using routine site
investigation data, allowing them to be used routinely in design and analysis. Soil-specific constitutive
parameters are derived from shearbox and oedometer test data, and are found to significantly
outperform existing empirical correlations based on relative density, highlighting the importance of
specifying a suitably detailed site investigation. An improved sliding-block (‘Newmark’) approach is
also developed for estimating permanent deformations during preliminary design, in which the
formulation of the yield acceleration is fully strain-dependent, incorporating both material hardening/
softening and geometric hardening (re-grading). The site-specific (improved) FE models and the new
sliding-block approach are shown to outperform considerably existing FE parameters and sliding-block
models in capturing the permanent deformations of the slope under virgin conditions, and further,
only the improved FE and sliding-block models are found to capture correctly the behaviour of the
‘damaged’ slope under subsequent earthquakes (e.g. strong aftershocks). The FE models can
additionally accurately replicate the settlement profile at the crest and quantify the dynamic motions
that would be input to supported structures, although these were generally overpredicted. The FE
procedures and sliding-block models are therefore complementary, the latter being useful for
preliminary design and the former for later detailed design and analysis.

KEYWORDS: centrifuge modelling; earthquakes; embankments; numerical modelling; sands; slopes

INTRODUCTION material strength characteristics, which may be either strain-


Transport infrastructure is a vital lifeline that must be safe- hardening or strain-softening (Matasovic et al., 1997; Wart-
guarded in seismic regions. Shallow embankment slopes and man et al., 2005), and to incorporate the dynamic behaviour
cuttings will commonly be used along the length of a road (e.g. ground motion amplification) within the sliding mass
or railway line to allow changes in gradient, and damage to itself in the case of deep rotational slips, as may be common
this type of infrastructure could inhibit the movement of in landfill slopes (Rathje & Bray, 1999; Wartman et al.,
emergency services and rebuilding in the aftermath of an 2003). Experimental work to validate such methods has been
earthquake. Although the infrastructure generally supported conducted using 1g shaking tables on cohesive soils (Wart-
on such constructions is relatively light (low bearing pres- man et al., 2005); however, there remains a need to demon-
sure) and flexible, significant damage can be caused by large strate the approach’s validity at realistic stress levels (e.g. by
permanent seismic slip within the slope. conducting centrifuge modelling). Although such analytical
A popular method of analysis for predicting seismic slip is tools are useful for parametric/comparative study, particularly
the Newmark sliding-block technique (Newmark, 1965). In in preliminary design, they are not able to predict the
this method a yield acceleration is calculated under pseudo- dynamic ground motions or settlement distribution (angular
static conditions using limit analysis techniques for simple distortion) behind the slope crest, both of which may have a
cases (e.g. Kim & Sitar, 2004), or numerical methods such controlling influence on the design of any infrastructure
as finite-element limit analysis (FELA) for more complex supported at the top of the slope. They are therefore not well
cases (e.g. Loukidis et al., 2003). A strong ground motion suited to detailed analysis or design.
(either a historically recorded motion or a synthetically Recent periods of seismic activity in Japan and New
produced accelerogram) is then used to determine the net Zealand (2011–2012) have demonstrated that civil engineer-
downslope acceleration, which can then be integrated nu- ing infrastructure may be subjected to several successive
merically to obtain slip velocity and slip displacement (Jib- strong ground motions within a short period (‘short’ here
son, 1993). Since it was originally proposed in the 1960s, meaning that there has been insufficient time to complete
this method has been substantially improved to incorporate remediation or reinstatement of damage caused by earlier
ground shaking). This means that there is a further need to
understand the behaviour of seismically damaged infrastruc-
Manuscript received 5 October 2012; revised manuscript accepted 7
June 2013. Published online ahead of print 15 July 2013.
ture under successive periods of ground shaking. This fea-
Discussion on this paper closes on 1 April 2014, for further details see ture is not currently incorporated in existing sliding-block
p. ii. methods.
 University of Dundee, UK, and Wasit University, Iraq. This paper will consider the behaviour of shallow cohe-
y sionless slopes under a sequence of strong earthquake
University of Dundee, Dundee, UK.

1230
AFTERSHOCKS AND THE WHOLE-LIFE SEISMIC PERFORMANCE OF GRANULAR SLOPES 1231
ground motions, and develop improved analysis tools for the (such as may be necessary during the early stages of design,
prediction of such behaviour. These tools will be applicable or in the immediate aftermath of an earthquake) or used
to the analysis of seismically damaged slopes under strong within a probabilistic, performance-based earthquake engin-
aftershocks, and to determination of the whole-life seismic eering framework if desired (Kramer, 2008). The FEM ap-
performance of slopes (i.e. under a range of successive proach can subsequently be used to undertake a more
motions of different strengths that the slope may see during detailed analysis of specific cases, which can incorporate the
its design life). dynamic behaviour of any infrastructure located at the slope
Fully dynamic numerical modelling in the time domain, crest.
using the finite-element method (FEM), will first be applied
to the problem, with the aim of producing a single analysis
in the time domain that captures both the dynamic vibration CENTRIFUGE MODELLING
effects and the permanent slope deformations. Existing non- Dynamic centrifuge testing was conducted using the 3.5 m
linear soil constitutive models will be used that encapsulate diameter beam centrifuge and servo-hydraulic earthquake
the strain history (‘seismic memory’) of the soil. Emphasis simulator at the University of Dundee. Two models were
will be placed on the efficient parameterisation of such a flown, representing identical slopes of 0  288 ( 1:2) at
model: (a) using previously published correlations based on 1:50 scale in dry sand at 50g. This slope angle was selected
databases of element test data (which require only relative to ensure that the soil was statically stable (0 , 9cs , the
density as an input); and (b) through the use of routine critical state friction angle of the soil), but with a suffi-
laboratory test data to produce improved soil-specific cali- ciently low factor of safety (and therefore low yield accel-
brations. The ability to model soil response realistically, eration, khy ) to ensure that large slip displacements would be
without requiring an excessive number of empirical param- generated during strong ground motion, such that the FEM
eters derived from non-standard tests, will allow the vali- and sliding-block models could be validated to large dis-
dated FEM procedures developed to be used with confidence placements. All subsequent dimensions and properties are
in routine geotechnical practice. given at prototype scale at 50g, unless otherwise stated.
An improved sliding-block method is also developed that The arrangement and instrumentation of the slope models
can fully capture the strain history of the soil, particularly is shown in Fig. 1(a). The slopes were prepared at a relative
by incorporating the effects of slope deformation (re-grad- density of ID ¼ 55–60% (the range accounts for the accu-
ing) during slip within the formulation of the yield accelera- racy in being able to measure and replicate ID ), were 8 m
tion. This allows the slip of a seismically damaged tall from toe to crest, and were underlain by a further 6 m
(deformed) slope to be predicted analytically under subse- of sand at the same relative density. HST95 (Congleton)
quent strong ground motion. silica sand was used, which has the basic properties given in
Both the FEM and sliding-block models are validated Table 1 (after Lauder, 2011). This uniformly rounded sand is
against centrifuge test data. The computationally simpler very fine, and has been used by other researchers at Dundee
sliding-block model, which requires no specialist software, to study other seismic phenomena (e.g Bertalot & Brennan,
can then be used to undertake rapid parametric analyses 2012). The values of emax and emin reported in Table 1 were

LVDTs

1 5
8 (160)

2 9 6
14 (280)

3 7 10 14 12
6 (120)

4 8 15 11 13

11 (220) 15 (300) 7·5 (150)

Accelerometers
(a)

Size of model in centrifuge – Fig. 1(a)

Positive acceleration
(downslope)

Absorbent boundary
(Lysmer & Kuhlmeyer, 1969)
(b)

Fig. 1. Slope configuration: (a) centrifuge model layout and location of ‘virtual’
instruments in numerical models; (b) finite-element mesh, showing boundary conditions
1232 AL-DEFAE, CAUCIS AND KNAPPETT

Magnitude: g/Hz
Table 1. State-independent physical properties of HST95 silica 150
sand (after Lauder, 2011) 100
50
Property Value
0
0 0·5 1·0 1·5 2·0 2·5 3·0 3·5 4·0 4·5 5·0
Specific gravity, Gs 2.63 Frequency: Hz
D10 : mm 0.09 (a)

Magnitude: g/Hz
D30 : mm 0.12 40
D60 : mm 0.17 30
Cu 1.9 20
Cz 1.06 10
Maximum void ratio, emax 0.769 0
0 0·5 1·0 1·5 2·0 2·5 3·0 3·5 4·0 4·5 5·0
Minimum void ratio, emin 0.467 Frequency: Hz
(b)

determined in accordance with BS1377 Part 4 (BSI, 1990). Fig. 3. Input bedrock motions in frequency domain: (a) Chi-Chi
The sand was pluviated in air using a slot pluviator into an (test AA01); (b) Kobe (test AA02)
equivalent shear beam (ESB) container having flexible walls,
the construction of which is described in Bertalot (2012).
This container was used to reduce potential boundary effects stronger shaking was directed in the downslope direction,
due to shear-wave reflection at the container walls. The which is represented by positive values of acceleration). The
model was instrumented within the soil by 15 type ADXL78 latter motion is well known to be particularly destructive to
MEMS accelerometers (70g range) manufactured by Ana- civil engineering infrastructure having a broad frequency
log Devices, and four external linear variable differential band below 3 Hz. The demand motions were bandpass-
transformers (LVDTs), with a pair of instruments measuring filtered between 0.8 Hz and 8 Hz (40–400 Hz at model
settlement at and behind the crest of the slope along the scale) using a zero-phase-shift digital filter to remove com-
centreline of the model, and a further pair placed adjacent ponents of the signal that were outside the range that can be
to one of the side walls to measure any boundary effects accurately controlled by the earthquake simulator. In both
(this latter pair will not be discussed further in this paper). tests, four nominally identical earthquake motions were
All settlements shown subsequently in this paper are taken applied to the model in succession, to investigate the behav-
from the rightmost instrument in Fig. 1(a). iour of the slope under strong aftershocks following initial
Ground motions were applied to the models using the strong shaking causing substantial slip.
Actidyn QS67-2 servo-hydraulic earthquake simulator re- Details of the model tests are summarised in Table 2. All
cently installed at the University of Dundee. The perform- ground motions were initially calibrated on a dummy model
ance of this actuator is described in Bertalot et al. (2012). identical to that shown in Fig. 1(a), but without instrumenta-
Earthquake ground motions were downloaded from the tion, to train the programmable logic controller within the
PEER (Pacific Earthquake Engineering Research) NGA data- earthquake simulator to achieve a faithful and repeatable
base. One of the models (test AA01) was subjected to a replication of the demand motions. As a result, the ground
horizontal strong ground motion recorded during the motions applied in each successive earthquake are felt to be
Mw ¼ 7.6 Chi-Chi earthquake in 1999 (Station TCU072, as close to identical as could realistically be achieved in
PGA ¼ 0.41g), and a second (test AA02) was subjected to a practice, this being an idealisation of the successive motions
horizontal motion recorded at the Nishi-Akashi recording having the same source (depth, faulting mechanism and
station in the Mw ¼ 6.9 Kobe earthquake in 1995 position).
(PGA ¼ 0.43g). These motions had approximately the same
peak acceleration, but different characteristics in the time
and frequency domains, as shown in Figs 2(a) and 2(b) and FINITE-ELEMENT MODELLING
Figs 3(a) and 3(b). Both records were recorded in ground The centrifuge tests were modelled using PLAXIS 2D in
with Vs . 450 m/s, representing shaking from stiffer layers plane strain with the mesh and boundary conditions shown
beneath the soil profile tested, such that any site amplifica- in Fig. 1(b). Compared with the centrifuge model shown in
tion occurs solely from the soil layer tested in the model. Fig. 1(a), the dimensions of the model domain were ex-
The former motion was used, as the Chi-Chi earthquake tended laterally and combined with non-reflecting boundary
caused a particularly high number of slope failures (Khazai elements controlling the dynamic stresses along the vertical
& Sitar, 2004), and is strongly directional (in these tests, the boundaries (after Lysmer & Kuhlmeyer, 1969) to represent
semi-infinite soil conditions, that is, boundary deformations
at the location of the centrifuge container wall that are
Acceleration: g

0·5 controlled by the dynamic deformation of the adjacent soil.


This boundary condition can also be modelled by horizontal
0
node-to-node ties between the two vertical boundaries of a
⫺0·5 model the width of the soil tested in the centrifuge. Com-
0 5 10 15 20 25 30 35 40 45 50 pared with this alternative, the method used has a higher
Time: s
(a) element requirement for the same mesh density, but allows
Acceleration: g

0·5

0 Table 2. Summary of centrifuge models tested


⫺0·5 Test ID : degrees ID : % Input motion (no.) Peak input
0 5 10 15 20 25 30 35 40 45 50
Time: s acceleration: g
(b)
AA01 28 56 Chi-Chi, 1999 (4) 0.41
Fig. 2. Input bedrock motions in time domain: (a) Chi-Chi (test AA02 28 59 Kobe, 1995 (4) 0.43
AA01); (b) Kobe (test AA02)
AFTERSHOCKS AND THE WHOLE-LIFE SEISMIC PERFORMANCE OF GRANULAR SLOPES 1233
information to subsequently be collected for points far from A reference pressure of pref ¼ 100 kPa is used throughout
the crest and toe of the slope (although these are not the remainder of this paper. Although the model has several
reported in this paper). A dynamic ground displacement was input parameters, all except Rf can be measured in some
applied along the bottom edge of the finite-element (FE) form through routine laboratory testing.
model, shown by the repeating arrows in Fig. 1(b). The In the first set of FEM simulations that will be described
input motion applied to the model was that measured at in the subsequent sections, the empirical correlations be-
instrument 8 in the centrifuge model: that is, the motion that tween the input parameters and relative density for coarse-
the slope in the centrifuge actually saw, accounting for any grained soils presented by Brinkgreve et al. (2010) were
losses between the shaking table and the container, and used. Use of these parameters required no additional soil
between the container bottom and the soil. The motions were testing, and all the constitutive parameters can be defined
input as ground displacements, determined by high-pass using relative density, ID , alone (which was uniform within
filtering and integration of the accelerometer records: filter- the centrifuge model slopes). Use of this model represented
ing before integration to obtain velocity, and again before the case in practice where no detailed laboratory test data
integrating velocity to obtain displacement, ensures that are available, but where relative density can be determined
there is no permanent ‘wander’ due to any offset in the from standard penetration tests (SPTs) or a cone penetration
accelerometer recordings or integration of random noise test (CPT) profile (e.g. as outlined in Knappett & Craig,
within the signal. Displacement data were extracted from the 2012).
FE models at the locations of the instruments in the A second set of simulations was then conducted in which
centrifuge tests shown in Fig. 1(a). At points 1–15, accelera- the key strength and stiffness parameters were calibrated for
tions were subsequently determined from double differentia- the soil used in the centrifuge tests using laboratory test
tion of the displacements. data, the stiffness parameters being important in modelling
the dynamic effects and shear-wave propagation, and the
strength properties controlling the permanent deformations.
CONSTITUTIVE MODELLING This model is henceforth termed ‘HST95’. A large amount
An elasto-plastic soil model with isotropic hardening of published shearbox test data was available from previous
(Schanz et al., 1999) is used in the modelling presented studies that used the HST95 sand (Lauder, 2011; Bransby et
herein, in which the elastic behaviour incorporates strain- al., 2011); this was used herein to derive soil-specific
dependent stiffness variation following the model proposed strength parameters, supplemented with oedometer tests to
by Hardin & Drnevich (1972), as modified by Santos & determine soil-specific stiffness parameters. In a practical
Correia (2001) case it would be necessary only to undertake tests appro-
G 1 priate for the in situ soil states (e.g. density in this case);
¼ (1) however, in this paper a range of test data were collated/
G0 1 þ 0:385js =s,0:7 j obtained for reconstituted samples covering a wide range of
relative densities such that a more complete model could be
where s is shear strain and s,0:7 is the shear strain at which developed for use in future simulations of physical test
G/G0 is 70%. Plastic failure is modelled using a cap-type results that use this material.
yield surface combined with the Mohr–Coulomb failure
criterion. This model is included in the PLAXIS FE suite
used for the work described herein as the ‘hardening soil Strength parameters
model with small-strain stiffness’ (Benz, 2006). Although Figure 4 shows a summary of shearbox data from a total
this model captures only strain-hardening (and not strain- of 38 tests conducted over a range of confining normal
softening) behaviour, it will subsequently be shown that the effective stresses between 5 kPa and 200 kPa, as summarised
effects of strain-softening on slope movement are minor in Table 3. Fig. 4(a) shows shear stress measurements at
when the earthquake is strong enough to induce significant critical state (when volumetric change had stopped), indi-
slip, as in the motions used in the centrifuge testing. The cating that the critical state friction angle is 9cs ¼ 328. Fig.
selection of appropriate soil strength (peak or critical state 4(b) shows the secant peak friction angles measured over the
friction angles) will be discussed in greater detail later in stress ranges considered (stress-independent values were
this section. determined by straight-line fits to the peak strength data, as
The model requires 13 input parameters: unit weights 9pk is stress-independent within the constitutive model). A
under saturated and dry conditions; three measurable effec- straight-line fit to the data as a function of relative density
tive stress strength parameters, 9pk , c9 and ł9 (angle of appeared to be appropriate; focusing on the data points
dilation), for use in the Mohr–Coulomb failure criterion; six below ID ¼ 80% gave
measurable stiffness parameters (which are stress dependent) 9pk ¼ 20I D þ 29 (degrees) (4)
describing the response to deviatoric loading (E50 ), compres-
sive loading (Eoed ), unload–reload cycles (Eur , ur ), small- Dilation angles are also shown in Fig. 4(b); the straight-
strain stiffness (G0 ) and a shear strain for describing the line fit for this data was found to be
shape of the G–s relationship (s,0:7 ); and finally, two ł9 ¼ 25I D  4 (degrees) (5)
empirical parameters Rf and m, the former controlling the
deviatoric stress at failure, and the latter controlling the These simple linear fits to 9pk and ł9 satisfy the dilatancy
variation of the stiffness parameters with effective confining relationship given by Bolton (1986), and predict the value of
stress. 9cs ¼ 328 with an error of , 1% using either this relation-
  ship, or that by Rowe (1962), which is incorporated in the
c9 cos 9   39 sin 9 m
E ð 39 Þ ¼ Eref (2) constitutive model formulation.
c9 cos 9 þ pref sin 9 Whereas the strength properties given by equations (4)
 
c9 cos 9  ð 39 =K 0 Þ sin 9 m and (5) match the element test data well, their use within a
Eoed ð 39 Þ ¼ Eref (3) strain-hardening model would imply that the peak strength is
oed
c9 cos 9 þ pref sin 9
appropriate to the analysis. It will subsequently be demon-
where E may be E50 , Eur or G0 ; Eref is the value of the strated that, for large-strain slope problems, the permanent
parameter at a reference stress of pref ; and K0 ¼ 1  sin 9. deformations are governed by the critical state strength.
1234 AL-DEFAE, CAUCIS AND KNAPPETT
140
Applied pressure
Shear stress at critical state, τ: kPa
CL
120

100

80

60 h ⫽ 19 mm
Lauder (2011)
40 Bransby et al. (2012)
This paper
20
φ⬘cs ⫽ 32°
0
0 50 100 150 200 250 Radius ⫽ 38 mm (⫽ 2h)
Normal effective stress, σ⬘n: kPa
(a)
Fig. 5. FE mesh used in simulating oedometer tests, showing
70 boundary conditions
Lauder (2011)
Peak friction angle, φ⬘pk; dilation angle, ψ⬘: degrees

Bransby et al. (2012)


60 were determined for loose and dense samples, from which
This paper
linear interpolations were made as a function of relative
50 Brinkgreve et al. (2010) density, as presented by Brinkgreve et al. (2010). The
φ⬘ interpolated parameters were subsequently checked against
40 further tests at intermediate densities. Fig. 6(a) shows a
comparison of some results for the Brinkgreve et al. (2010)
30 model, which overpredicts stiffness in dense sand and dra-
matically underpredicts stiffness in loose sand. Fig. 6(b)
20 shows the markedly improved results using fitted values of
ψ⬘ Eoed , E50 , Eur and m. To reduce the number of independent
10 parameters, E50 ¼ 1.25Eoed and Eur ¼ 3Eoed were assumed,

0
Applied veritical effective stress: kPa

0 20 40 60 80 100 700
⫺10 Relative density, ID: %
(b) 600
500
Fig. 4. DSA (shearbox) test data used in soil-specific calibration of
constitutive model: (a) 9cs ; (b) 9pk and ł9 400 Oedometer, ID ⫽ 20%
300 FEM, ID ⫽ 20%
Hence, in modelling the centrifuge tests, 9pk ¼ 9cs ¼ 328 200 Oedometer, ID ⫽ 73%
and ł9 ¼ 08 were used. FEM, ID ⫽ 73%
100
0
Stiffness parameters 0 1 2 3 4 5
Vertical strain: %
Applied veritical effective stress: kPa

Oedometer tests were conducted on dry samples of sand (a)


700
prepared by air pluviation within a Clockhouse Engineering
Ltd J550 oedometer (sample height ¼ 19 mm; sample diam- 600
eter ¼ 76 mm) at a range of relative densities, ID , between 500
5% and 83%. The tests were conducted up to an effective 400 Oedometer, ID ⫽ 20%
stress of 600 kPa, and included three unload–reload cycles FEM, ID ⫽ 20%
300
(200–100–200 kPa; 400–200–400 kPa; 600–400–600 kPa)
200 Oedometer, ID ⫽ 73%
to ensure that Eur was well calibrated (this parameter is
likely to be important during cyclic seismic loading). The 100 FEM, ID ⫽ 73%
constitutive model parameters were then determined by con- 0
ducting virtual oedometer tests using FEM, starting with the 0 1 2 3 4 5
parameters suggested by Brinkgreve et al. (2010), and Vertical strain: %
(b)
modifying them as necessary to achieve a reasonable fit to
the data. Fig. 5 shows the axisymmetric model geometry Fig. 6. Comparison of one-dimensional compression curves for
and boundary conditions employed. To develop a complete loose and dense samples: (a) using Brinkgreve et al. (2010)
density-dependent model in this paper, calibrated parameters parameters; (b) using HST95 (soil-specific) parameters

Table 3. DSA test data for HST95 silica sand

Source ID : % No. of tests Effective normal stresses: kPa

Lauder (2011) 17 8 5, 8, 11, 16, 30, 35, 70, 125


40 6 16, 30, 55, 100, 135, 150
75 4 11, 16, 30, 70
Bransby et al. (2011) 9 5 10, 25, 50, 100, 200
41 5 15, 35, 55, 100, 150
93 5 10, 25, 50, 100, 180
This paper (see Fig. 17) 55 5 5, 8, 13, 16, 25
AFTERSHOCKS AND THE WHOLE-LIFE SEISMIC PERFORMANCE OF GRANULAR SLOPES 1235
so that Eoed could be used to simulate the strain magnitude Chi-Chi (centrifuge data)
correctly, and m used to control the shape of the stress– Kobe (centrifuge data)
strain curve. E50 ¼ Eoed was proposed by Brinkgreve et al. Chi-Chi (FEM data)
Kobe (FEM data)
(2010), but this is not exactly true from comparison of
Hardin & Drnevich (1972)
equations (2) and (3). The value of 1.25 ensured that no
Ishibashi & Zhang (1993)
unrealistic values of K0 would be implied in denser soils Santos & Correia (2001)
(including that used in the centrifuge tests); the analyses 1·0
appeared to be relatively insensitive to this assumption at
lower densities. Following iteration, the best-fit stiffness
0·8
parameters were determined to be
Eref :
oed ¼ 25I D þ 20 22 (MPa) (6) 0·6

G/G0
: :
m ¼ 0 6  0 1I D (7)
0·4
The power coefficient m is the inclination of stiffness–
stress curves normalised by pref in double logarithmic scale. 0·2
The proposed calibration for m (equation (7)) gives a power-
law exponent for stress dependence of stiffness that is be-
tween 0.5 and 0.6 at all densities. This is consistent with 0
0·001 0·01 0·1 1
previous studies (e.g. Lo Presti et al., 1998), while maintain- Shear strain: %
ing the slight negative correlation between the two param- (a)
eters noted by Brinkgreve et al. (2010).
As neither of the aforementioned tests measures small- Chi-Chi (centrifuge data)
strain parameters, G0 was estimated using the relationship Kobe (centrifuge data)
based on void ratio (e) proposed by Hardin & Drnevich Chi-Chi (FEM data)
(1972), Kobe (FEM data)
0·5 Hardin & Drnevich (1972)
(2:97  e)2
Gref
0 ¼ 33 (MPa) (8) Ishibashi & Zhang (1993)
1þe 0·4
Santos & Correia (2001)

for pref ¼ 100 kPa. Equation (8) can be linearised, ignoring


Damping

0·3
terms of order e 2 and above; expressing e in terms of
relative density with the values of emax and emin gives
0·2
Gref :
0 ¼ 50I D þ 88 80 (MPa) (9)
0·1
The shear strain parameter s,0:7 was assumed to increase
linearly from 0.01% at ID ¼ 20% to 0.02% at ID ¼ 80%: that 0
is 0·001 0·01 0·1 1
Shear strain: %
s,0:7 ¼ 1:7I D þ 0:67 (310 )
4
(10) (b)

Fig. 7. Comparison of shear modulus degradation and damping


in centrifuge tests and FE simulations, along with other published
Other parameters and comments correlations
Of the remaining parameters, the default value of Rf ¼ 0.9
was used, and the unit weight parameters were determined proposed by Brennan et al. (2005). The data points for the
from standard relationships as a function of relative density numerical simulations were determined in the same way,
and linearised, giving using data from ‘virtual’ accelerometers at homologous
points within the FE mesh. Fig. 7 demonstrates that the
ªdry ¼ 3I D þ 14:5 (kN=m3 ) (11) operative shear modulus at a given cyclic shear strain is
ªsat ¼ 1:8I D þ 18:8 (kN=m )3
(12) comparable to that measured in the centrifuge tests, but that
the material hysteretic damping within the model is gener-
The particular choice of direct shear apparatus (DSA) and ally smaller than that inferred from the centrifuge tests. This
oedometer tests for determining the parameters, as described will be revisited in the following section. Nonetheless, it
above, was guided by available test data. Triaxial test data would appear that the use of simple monotonic test data is
could equally be used to determine both strength and effective in defining a constitutive model that can represent
stiffness data, by simulating triaxial compression on a virtual the principal dynamic behaviour of the soil.
test sample using FEM in a similar way to that described for
the oedometer tests. Damping will be discussed later in the
paper. Accelerations
Figure 8 shows a comparison of the measured and simu-
lated acceleration at the toe of the slope (instrument 14) in
VALIDATION OF FEM the first earthquake of test AA01 (Chi-Chi motion) in both
Dynamic shear modulus and damping the time and frequency domains (Figs 8(a) and 8(b) respec-
Shear modulus and damping as functions of cyclic shear tively). In this figure three cases are considered: (i) the use
strain within both the centrifuge models and the numerical of Brinkgreve et al. (2010) constitutive parameters; (ii) the
simulations are shown in Fig. 7. The data points representing use of HST95 parameters; and (iii) the use of HST95
the centrifuge models were determined from second-order parameters with additional Rayleigh damping. The constitu-
estimates using the accelerometer data, following the method tive model implicitly includes material hysteretic damping
1236 AL-DEFAE, CAUCIS AND KNAPPETT
Centrifuge Centrifuge

Acceleration: g
Brinkgreve et al. (2010) FEM
Acceleration: g
Case (i) Brinkgreve et al. (2010) FEM Case (i)
0·5 0·5
0 0
⫺0·5 ⫺0·5
0 5 10 15 20 25 30 35 40 45 50 0 5 10 15 20 25 30 35 40 45 50

Centrifuge Centrifuge

Acceleration: g
Acceleration: g

Case (ii) HST95 silica sand FEM Case (ii) HST95 silica sand FEM
0·5 0·5
0 0
⫺0·5 ⫺0·5
0 5 10 15 20 25 30 35 40 45 50 0 5 10 15 20 25 30 35 40 45 50

Centrifuge

Acceleration: g
Centrifuge
Acceleration: g

Case (iii) Case (iii) HST95 sand ⫹ Rayleigh FEM


0·5 HST95 sand ⫹ Rayleigh FEM 0·5
0 0
⫺0·5 ⫺0·5
0 5 10 15 20 25 30 35 40 45 50 0 5 10 15 20 25 30 35 40 45 50
Time: s Time: s
(a) (a)

Centrifuge

Magnitude: g/Hz
Magnitude: g/Hz

40 40 Brinkgreve et al. (2010) FEM


Case (i) Centrifuge
30 Brinkgreve et al. (2010) FEM 30 Case (i)
20 20
10 10
0 0
1 2 3 4 5 1 2 3 4 5
Magnitude: g/Hz

40 Case (ii) Centrifuge Centrifuge


Magnitude: g/Hz

HST95 silica sand FEM 40 HST95 silica sand FEM


20 Case (ii)
20
0
1 2 3 4 5 0
1 2 3 4 5
Magnitude: g/Hz

40 Centrifuge
Case (iii)
HST95 sand ⫹ Rayleigh FEM Centrifuge
Magnitude: g/Hz

20 40 HST95 sand ⫹ Rayleigh FEM


Case (iii)
0 20
1 2 3 4 5
Frequency: Hz
(b) 0
1 2 3 4 5
Frequency: Hz
Fig. 8. Comparison of measured and predicted accelerations at (b)
slope toe during test AA01: (a) time domain; (b) frequency
domain Fig. 9. Comparison of measured and predicted accelerations at
slope crest during test AA01: (a) time domain; (b) frequency
domain
within in its formulation; however, these models generally
overpredicted accelerations (as a result of underpredicting
damping; see Fig. 7). The Rayleigh damping formulation nificant interest for being able to determine the input motion
allows additional mass and/or stiffness-proportional (modal, to any (infra)structure located at the crest of the slope. For
frequency-dependent) damping (add ) to be included both cases (i) and (ii) without any additional damping,
  significant overprediction of acceleration is observed. From
1
add ¼ cm þ ck ( f n ) (13) Fig. 9(b) it is clear that this appears to arise from increased
4 f n amplification of the higher-frequency components (above
3 Hz). The Rayleigh damping parameters reported earlier
where add is the additional equivalent viscous damping give damping that is predominantly stiffness-proportional,
ratio, and fn is the natural frequency of modes within the such that the motions at higher frequencies would be more
soil. Values of cm ¼ 0.0005 and ck ¼ 0.005 were selected for significantly damped, without over-damping the lower fre-
use in case (iii), and the justification for these will be given quencies where the match is better. Case (iii) in Fig. 9(b)
later. shows a markedly improved prediction of the acceleration
At point 14 the match between the centrifuge test and time history at this point, owing to reduced contribution of
FEM is good for all soil parameter sets considered, although the higher-frequency modes. Similar results were found for
the models without the additional Rayleigh damping slightly simulations of test AA02 (Kobe motion). Increases in the ck
overpredict the peak acceleration values in some of the parameter, to increase and further improve the damping at
cycles. This match is not surprising, given that the shear higher frequencies, resulted in poorer predictions of perma-
wave has propagated only a short distance in level ground at nent movements. The values for cm and ck given above
this point, and has not yet interacted with the slope. appear to represent the best compromise, enabling both
Figure 9 shows similar plots at the crest of the slope ground accelerations and permanent slip (see later) to be
(instrument 5), representing a more stringent test of the reasonably estimated within the same analysis.
capabilities of the numerical model. This would be of sig- The ratios of crest peak acceleration (at instrument 5) to
AFTERSHOCKS AND THE WHOLE-LIFE SEISMIC PERFORMANCE OF GRANULAR SLOPES 1237
the peak value of the input motion were found to be comparison of predicted crest spectra for 5% nominal struc-
insensitive to the number of earthquakes, and of the order of tural damping from FEM and measured data from the
1.3 (highest value recorded was 1.4). This amplification centrifuge tests, also including design spectra based on
factor clearly contains two distinct effects: the site effect (in Eurocode 8 for context. For the cases including Rayleigh
essence a material property effect, due to the dynamic damping (HST95, case (iii)), a very good match to the
properties of the soil); and a topographic effect (a geometric centrifuge data is obtained, although there is a tendency for
effect due to the ground surface profile). In Eurocode 8, Part the response to be slightly underpredicted for periods above
5 (BSI, 2005b), the ground motion at the crest (instrument 0.5 s. Use of the other constitutive models results in a much
5), for use in constructing response spectra, is calculated more significant underprediction of response above 0.5 s,
using and overprediction below this. Fig. 10 also suggests that
there is a range of natural periods (between approximately
PGA ¼ S  S T  ag (14)
0.4 s and 1.0 s) over which the FEM may substantially
where ag is the peak acceleration in the underlying bedrock underpredict the response, compared with the centrifuge data
(input motion in the centrifuge tests); S is the ‘soil factor’ (this is particularly noticeable for test AA02). This range
describing the site effect (¼ 1.4 for the ‘ground type E’ soil could be used as a simple screening tool to identify key
in this study, as classified using Eurocode 8, Part 1 (BSI, pieces of infrastructure atop slopes that may be more vulner-
2005a)); and ST is the topographic amplification factor able to seismic damage than FEM would suggest, and to
(> 1.2 for shallow slopes). The overall amplification factor which extra consideration should be paid in design. It is
is then PGA/ag : Based on Eurocode 8, the overall amplifica- clear that care should be taken in interpreting the hazard to
tion within the centrifuge test would be predicted to be supported infrastructure from analyses using FEM, and that
between 1.4 and 1.7, which would be conservative when this will be significantly dependent on the natural period of
compared with the centrifuge observations (amplification  the supported structure. Further comparison between the first
1.3) for the two motions considered. Ashford et al. (1997) and last earthquake motions demonstrated that the spectral
and Bouckovalas & Papadimitriou (2005) have conducted magnitudes appear to be relatively insensitive to repeated
detailed numerical studies based on harmonic ground shak- strong earthquake shaking.
ing, and have demonstrated that the topographic amplifica-
tion is dependent on the ratio of slope height to wavelength
and position from the crest. In light of this, the centrifuge Permanent deformations
tests of Brennan & Madabhushi (2009) are a useful compari- Figure 11 shows the permanent crest settlements across
son with the test data in this paper, as they considered a the four earthquakes as predicted by FEM (all three cases
slope of similar (although not identical) height, slope angle are shown), and as measured in the centrifuge. It is clear
and sand. Their results support the value of 1.3 measured in that, for both suites of earthquake motions, use of the
tests AA01 and AA02. Brinkgreve et al. (2010) parameters hugely overpredicts
However, the hazard posed to infrastructure at the crest is settlement at the crest. This would lead to significant over-
more usefully represented by a response spectrum, rather prediction of the risk posed to the slope, and hence, poten-
than a ratio of peak accelerations, which represents the tially, uneconomic design. In contrast, use of the HST95
spectral response at a period T ¼ 0 s only. Fig. 10 shows a parameters (either case) gives a much better prediction,
although inclusion of the additional damping, which was
beneficial for modelling the dynamic behaviour accurately,
results in an underprediction of permanent deformation
3·5 Centrifuge (AA01) (which is unsafe). The Brinkgreve et al. (2010) parameters
Brinkgreve et al. (2010) FEM
3·0 were initially envisaged as being useful for cases where
Spectral acceleration: g

HST95 sand FEM


2·5 HST95 sand FEM ⫹ Rayleigh there is extremely limited site investigation data, but the
EC8 (S ⫽ 1·4, ST ⫽ 1·2) implication of Fig. 11 is that it is always important to obtain
2·0 soil parameters from high-quality laboratory (or in situ) test
1·5 data to achieve an accurate prediction of movement. It is
1·0
likely that the extra cost of this additional investigation
would be significantly offset if a lower amount of remedia-
0·5 tion/repair were to be required (as a result of a more
0 accurate prediction of seismic response). When considered
0 0·2 0·4 0·6 0·8 1·0 1·2 1·4 1·6 1·8 2·0
Period: s
alongside the dynamic performance discussed in the previous
(a) section, it is clear that neither case (ii) or case (iii) can give
consistently better performance of both dynamic and perma-
3·5
Centrifuge (AA02) nent movements simultaneously; this would appear to sug-
3·0 Brinkgreve et al. (2010) FEM gest that the additional Rayleigh damping may not be a
Spectral acceleration: g

HST95 sand FEM material characteristic, but may be masking an effect of the
2·5 HST95 sand FEM ⫹ Rayleigh sloping ground geometry in which wave reflection at the
EC8 (S ⫽ 1·4, ST ⫽ 1·2)
2·0 sloping ground surface is not modelled correctly within the
1·5 FE model.
Figure 12 shows a comparison (drawn to scale) of the
1·0
ground surface profile measured at the end of the centrifuge
0·5 test (a negligible amount of movement was recorded during
0
spin-down), and as predicted at the end of the last earth-
0 0·2 0·4 0·6 0·8 1·0 1·2 1·4 1·6 1·8 2·0 quake (EQ4) from the FE model (case (ii)) for test AA01. It
Period: s can be seen that the numerical model captures the deformed
(b)
shape of the slope well, particularly the angular distortion at
Fig. 10. Measured (centrifuge), predicted (FEM) and design the crest, which is likely to be of greatest significance for
(Eurocode 8) response spectra at top of slope (instrument 5) for supported infrastructure. A similar result was obtained for
5% structural damping: (a) Chi-Chi (AA01); (b) Kobe (AA02) test AA02 (not shown).
1238 AL-DEFAE, CAUCIS AND KNAPPETT
0

0·5

1·0

1·5

2·0

Settlement: m
2·5

3·0

3·5

4·0 Centrifuge settlement (0·91 m)


Brinkgreve et al. (2010) FEM (5·05 m)
4·5
HST95 sand FEM (0·79 m)
5·0
HST95 sand FEM ⫹ Rayleigh (0·41 m)
5·5
0 50 100 150 200
Time: s
(a)
0

0·5

1·0
Settlement: m

1·5

2·0

2·5 Centrifuge settlement (1·02 m)


Brinkgreve et al. (2010) FEM (3·3 m)
3·0 HST95 sand FEM (0·76 m)
HST95 sand FEM ⫹ Rayleigh (0·32 m)
3·5 0 50 100 150 200
Time: s
(b)

Fig. 11. Comparison of permanent crest settlements from FEM and centrifuge modelling:
(a) test AA01; (b) test AA02
0·79 m (HST95 FEM)

β
0·91 m (centrifuge)

khg L
Original profile
β g
khγzL cos β

After centrifuge
8 (160)

τultL
After FEM z
14 (280)

(σ⬘ ⫹ u)L
γzL cos β
6 (120)

(Centrifuge) 1·95 m

Fig. 13. Forces acting within infinite slope under horizontal


shaking
11 (220) 15 (300) 7·5 (150)

Fig. 12. Comparison of slope profile after EQ4 as predicted by uniaxial horizontal shaking (plane strain), the applied down-
FEM and as measured in centrifuge, test AA01 slope shear stress is
applied ¼ ªz sin  cos  þ k h ªz cos2  (15)
DEVELOPMENT OF AN IMPROVED SLIDING-BLOCK
METHOD where the first term relates to the static shear stress due to
The horizontal yield acceleration of a shallow translational the ground slope, and the second term relates to the addi-
(infinite) slip can be determined using standard limit equi- tional peak dynamic shear stress induced by the earthquake.
librium techniques, incorporating pseudo-static acceleration The shear strength of the soil along the slip plane, assuming
components due to the seismic ground motion (see Fig. 13). that the soil failure can be described by the Mohr–Coulomb
For a slip plane at depth z beneath the slope surface under failure criterion, is given by
AFTERSHOCKS AND THE WHOLE-LIFE SEISMIC PERFORMANCE OF GRANULAR SLOPES 1239
ult ¼ c9 þ  9 tan 9 the slip plane (so long as it continues to be parallel to the
(16) slope surface). Equation (17) then simplifies to
¼ c9 þ (ªz cos2   k h ªz sin  cos   u) tan 9
cos  tan 9  sin 
k hy ¼ (18)
The soil yields when applied ¼ ult : The value of kh at cos  þ sin  tan 9
which this occurs (the yield acceleration, khy ) can be deter-
mined from equations (15) and (16) as In equation (18) the only parameters affecting khy are the
slope angle  (geometric) and the soil friction angle 9
c9 þ (ªz cos   u) tan 9  ªz sin  cos 
2
(constitutive). In a standard analysis, both  and 9 are
k hy ¼ (17)
ªz cos2  þ ªz sin  cos  tan 9 constant. In reality, however, the soil may be strain-softening,
in which case 9 will depend on the magnitude of the shear
In a standard Newmark analysis, when the horizontal strain (s ) on the slip plane and the density of the soil.
component of the ground acceleration (a) exceeds khy in the Matasovic et al. (1997) presented a simplified model for this,
downslope direction, the slope will start to accelerate under which is shown schematically in Fig. 14(b). If 9pk ¼ 9cs , the
the slip acceleration, aslip ¼ a  khy (positive value implies model reduces to the standard case of a strain-softening
downslope movement). This acceleration is numerically in- material. To incorporate this into an analysis, the value of
tegrated with respect to time to obtain the slip velocity, khy at a particular time step can be estimated, based on the
which is then itself integrated to obtain the slip displace- current permanent downslope displacement, computed in the
ment. Once a , khy , the sliding block will begin to decele- previous time step, divided by the thickness of the shear
rate (as aslip , 0) until the slip velocity reaches zero, at band/slip plane to obtain an estimate of the shear strain.
which point the block comes to rest until aslip is again The slope angle  will also change during an analysis, as
positive. This procedure is shown schematically in Fig. slip will cause settlement at the crest and accumulation of
14(a). material at the toe: that is, the slope will become shallower
In a dry cohesionless soil, c9 ¼ 0, u ¼ 0, and ªz cancels (re-grading, RG). A simplified model for this geometric
in equation (17): that is, khy is independent of the depth of effect is developed in this paper, and is shown in Fig. 15.
Fig. 15(a) shows the kinematically admissible failure me-
chanism assumed for an increment of slip, d i , in which
Including strain-softening (SS): Classical Newmark: infinite sliding is the predominant component. This leads to
khy ⫽ f (φ⬘) khy ⫽ const. downward vertical movement of the material at the slope
crest and a horizontal translation of the position of the toe,
Acceleration

resulting in a new, shallower slope with an approximate


slope angle of  iþ1 in Fig. 15(b). Ambrayses & Srbulov
(1995) present an alternative mechanism for shallow sliding
Time that is similar to Fig. 15(a), but with more complex changes
in geometry (and for post-seismic sliding only, i.e. when
Ground motion there is no seismic inertia). The new (simpler) mechanism
proposed here has the advantage that the calculations are
much more straightforward, relying on only a single param-
eter (the initial slope angle ) to define the slope geometry,
Slip velocity

Classical

Including SS
and more closely match the deformations observed visually
in the centrifuge tests (e.g. see Fig. 12).
Provided that  is relatively small (such that the slope is
Time
long compared with its height), the equilibrium of this
mechanism will be well approximated by infinite slope
theory (i.e. equations (17) or (18) will adequately describe
Slip displacement

Crest Toe
Settlement increment
di sin βi

Slip increment,
di
Time
(a)

di cos βi
Including strain-softening (SS):
φ⬘pk khy ⫽ f (φ⬘) (a)
New slope
surface (i ⫹ 1)
φ⬘cs
khy and φ⬘

Hi βi⫹1
Classic Newmark: Hi ⫺ di sin βi
Shearbox test curve
khy ⫽ const. tan⫺1 (τ/σ⬘) vs εs
βi

εs,pk εs,cs Shear strain, εs Hi cot βi di cos βi


(b) (b)

Fig. 14. Newmark sliding-block procedure, and effect of strain- Fig. 15. (a) New incremental slope re-grading mechanism;
softening (after Matasovic et al., 1997) (b) incremental changes in geometry
1240 AL-DEFAE, CAUCIS AND KNAPPETT
khy ). It will be demonstrated later that even a slope as steep
as ,1:2 satisfies this condition to a reasonable degree of
accuracy. It is assumed that any volumetric change in the 0·5 m
material in the sliding block is negligible. From Fig. 15(b),
the instantaneous slope angle  iþ1 can be determined from
  0 12345678
H i  d i sin i
iþ1 ¼ tan 1
(19) (a)
H i cot i þ d i cos i
Shear strain: %
400
where Hi is the height of the slope at the previous time step. 360
For the initial time step, d0 ¼ 0, Hi ¼ H and  i ¼ 0 (initial 320
slope angle). It should be noted that the deformations in Fig. 280
240
15(b) are shown at exaggerated scale for clarity. The slope 200
angle can therefore be recalculated at each time step to 160
account for the re-grading of the slope based on the 120
increment of slip occurring in the previous time step, as 9 80
40
was previously to account for strain-softening. The yield 0
acceleration (equation (18)) will thus be fully strain-depen-
(b)
dent. This method assumes that once the slope has been
deformed to a new, smaller value of , the failure mechan- Fig. 16. (a) Failure mechanism computed from DLO for seismic
ism will continue to be of the infinite type, with a new slip case; (b) accumulated shear strain within FE model (test AA01)
surface forming parallel to the new slope surface. It also and assumed approximate infinite slip mechanisms used in
assumes that the strain-dependent effects on  and 9 are Newmark analyses
independent, to simplify the calculations. In reality, this
latter assumption may not be wholly true, as the changing
angle of a softened slip plane (i.e. with 9 ¼ 9cs ) may push of effective confining stresses representing those within the
at least part of the slip surface into previously undisturbed top 1 m of the soil. The test data are shown in Fig. 17, from
soil. If this effect and the effects of strain-softening are which values of 9pk ¼ 408, 9cs ¼ 328, s;pk ¼ 3.5% and
significant, it is expected that the model will overestimate s;cs ¼ 7.5% were estimated for 0.5 m depth. Using these
movements; however, this would be conservative for use in friction angles and limiting shear strains, the initial yield
analysis and design. It should also be noted that the model accelerations of the slope were computed using equation
as formulated can be used even for the case of large total (18) and DLO; the results are shown in Table 4, alongside
slope movements (such as may accrue during a series of values for the static factor of safety of the slope (F),
strong aftershocks), as the displacement increment in each calculated using equations (15) and (16) with kh ¼ 0, and
individual time step will remain small, and therefore the DLO.
instantaneous failure mechanism will be represented by Fig. The shear band thickness (required for converting slip
15(a) for small displacement increments. displacement in the sliding-block model into an approximate
From the form of equation (19) it is clear that  can only shear strain) was estimated at t ¼ 16D50 ¼ 2.4 mm, based on
ever reduce during an earthquake, resulting in an increase of a range of previous studies (e.g. Mülhaus & Vardoulakis,
khy from equation (17) or (18): that is, the slope will 1987; Oda & Kazama, 1998; Muir Wood, 2002). The
geometrically harden during an earthquake, and the slip in a calculations were conducted at prototype scale, and so the
subsequent (identical) earthquake will be less than that shear strain was estimated using 50t ¼ 120 mm to model the
occurring in the original. The behaviour of a seismically correct ratio between the slope geometry and the grain size
damaged slope during a subsequent earthquake can therefore within the model. For application to a true field case where
be determined by starting from the initial conditions (amount the grains are smaller compared with the overall size of the
of slip, accumulated strain, re-graded slope angle and current slope, the true shear band thickness should be used instead.
friction angle) obtained at the end of the analysis for the For the tests presented herein, changing the shear band
previous ground motion. thickness from 120 mm to 2.4 mm resulted in less than 1%
change in crest settlement (the actual value varied slightly
with the input motion considered), confirming that the grain-
VALIDATION OF SLIDING-BLOCK MODEL
Predictions of permanent deformation 1·0
Visual observations from the centrifuge tests suggested
that the 1:2 slopes tested failed in a shallow translational
mechanism consistent with that shown in Fig. 15. The depth 0·8
Stress ratio (⫽ tan φ⬘)

of the shear plane ( 0.5 m) was estimated using the dis-


continuity layout optimisation (DLO) technique (Smith & 0·6
Gilbert, 2007) to obtain a minimum upper-bound mechanism
for the actual limited geometry of the centrifuge model.
0·4 5·5 kPa (0·34 m)
DLO calculations were carried out using LimitState:GEO,
v2.0. The position of the slip plane was not affected by the 8·2 kPa (0·51 m)
value of 9 used, and is shown in Fig. 16(a). The idealised 0·2 13·6 kPa (0·85 m)
geometry assumed in the analysis (cf. Fig. 15(a)) is shown
Analytical model
in Fig. 16(b), overlaid onto the accumulated shear strain
within the FE simulation of test AA01 (case (ii)), for 0
0 5 10 15 20
comparison. Shearbox testing was then conducted on sam- Shear strain, εs: %
ples of dry sand prepared to the same relative density as in
the centrifuge tests, using a standard 60 mm 3 60 mm DSA, Fig. 17. Soil test data from direct shear apparatus (DSA) at low
to obtain 9pk and 9cs : These tests were conducted at a range effective confining stress
AFTERSHOCKS AND THE WHOLE-LIFE SEISMIC PERFORMANCE OF GRANULAR SLOPES 1241
Table 4. Static and dynamic slope stability data

Soil strength Static stability parameters Dynamic stability parameters

F (equations (15) F (DLO) khy khy (DLO)


and (16)) (equation (18))

9 ¼ 328 1.17 1.20 0.07g 0.07g


9 ¼ 408 1.56 1.61 0.21g 0.22g

size scaling effect is negligible, and that centrifuge model- behaviour of the material within the sliding block (due to its
ling is therefore an appropriate technique for modelling thickness) may be significant (e.g. Rathje & Bray, 1999);
slope failure problems in coarse-grained soil. A similar con- this was not the case for the extremely shallow translational
clusion was reached by Anastasopoulos et al. (2007) for slips that occurred within the cohesionless slopes tested
fault rupture (shear band) propagation through the same herein.
sand. Figures 19 and 20 show the results of simulations of
Figure 18 shows the effect of the geometric re-grading cumulative crest displacement both with (SS + RG) and
(change in ), using the first earthquake (EQ1) of test AA01 without re-grading (SS) with the centrifuge test data, for
as an example. Only the positive (downslope) accelerations tests AA01 and AA02 respectively. It can be seen that, in
have been shown, for clarity. No slip occurs until the ground each case, the improved model presented in this paper
motion exceeds the yield acceleration based on the peak (SS + RG) tracks the settlement at the crest of the slope
strength. Once the ground motion exceeds this value, how- much more closely than the model that incorporates only the
ever, and the slope begins to slip, the shear strain rapidly constitutive effect (SS). These latter models increasingly
accrues, resulting in softening to critical state conditions diverge from the measured values with further strong shak-
after the first large pulse. Motion of the slope also causes ing, as they always start with the initial (steeper) slope
re-grading (geometric hardening), and the yield acceleration geometry, and therefore overpredict the slip. If the input
can subsequently be seen to increase non-linearly throughout motions were identical, the slip in each subsequent earth-
the remainder of the earthquake, leading to reduced slip quake would be identical for the case of no re-grading
velocities (and hence reduced permanent slip) compared (although the movements in the first earthquake might be
with the case with no geometric hardening. slightly smaller, owing to the strain-softening effect). The
Sliding-block analyses were subsequently conducted for improved models are not perfect, and in each case over-
each of the centrifuge tests (for all four earthquakes in the predict the measured movement; as the slope re-grades, the
cases of tests AA01 and AA02). Simulations were conducted new position of the slip plane may cause it to pass at least
using the ‘bedrock’ input motion (this was taken from the
bottom-most accelerometer in the model, instrument 8), as
used in the FEM. Additional simulations using instrument 6 Time: s
did not show significant differences in the overall cumulative 0 50 100 150 200
slip predicted, possibly as there was some shear decoupling 0
across/around the shear band, which counteracted (at least
partially) any increase in acceleration due to soil amplifica-
Crest settlement: m

tion. In slopes with deep rotational failure surfaces (e.g. in


municipal solid waste or steep cohesive slopes), the dynamic
2
Centrifuge (AA01)
0·40 Sliding block (SS ⫹ RG)
Input ground motion Sliding block (SS)
Strain- khy ⫺ SS ⫹ RG
0·30
Acceleration: g

softening EQ1 EQ2 EQ3 EQ4


khy ⫺ SS
4
Geometric hardening
0·20 (re-grading)
0·5
0·10
0·21 0·23 0·25
Input acceleration: g

0·20
0 0·16
0 10 20 30 40 50
Time: s
(a) 0
0 50 100 150 200
1·0 Strain-softening and re-grading (SS ⫹ RG)
Strain-softening only (SS)
Slip velocity: m/s

0·8

0·6
⫺0·5
0·4

0·2
Ground acceleration (AA01)
0 khy (SS ⫹ RG)
0 10 20 30 40 50 khy (SS)
Time: s
(b)
Fig. 19. Comparison of predicted cumulative crest settlements
Fig. 18. Application of new sliding-block model, showing key (with and without re-grading) with centrifuge test measurements:
features (Chi-Chi EQ1 used, test AA01) test AA01 (Chi-Chi)
1242 AL-DEFAE, CAUCIS AND KNAPPETT
Time: s
0 50 100 150 200
strain-dependent sliding-block model (strain-softening and
0 geometric hardening) and the laboratory-test-calibrated FEM
developed in this paper (HST95, case (ii)). Although the
improved models are not perfect, they give a much im-
Crest settlement: m

proved prediction of the response under the initial earth-


quake on virgin soil, and are subsequently able to give a
2 better prediction of behaviour, even after several previous
Centrifuge (AA02) strong earthquakes during which significant deformation has
Sliding block (SS ⫹ RG) accrued. Both existing models overpredict settlements in the
Sliding block (SS)
first earthquake and then get progressively worse with
EQ1 EQ2 EQ3 EQ4 further shaking, as they are unable to correctly capture the
4 effect of the previous seismic (strain) history on the slope.
In the case of the sliding-block model this is associated
0·5 with incorrect description of the deformed slope; in the
case of the FEM, the case (i) constitutive parameters do
0·21 not appear to correctly capture the strain history or mech-
Input acceleration: g

0·18 0·20
0·12 0·16 anical response of the soil. The development of the ‘im-
proved’ sliding-block and FEM tools in this paper provides
0 an improved means of quantifying the response of shallow
0 50 100 150 200 cohesionless slopes under strong earthquake shaking, and
the ability to consider behaviour under multiple successive
earthquakes. With further development, these tools will
allow civil engineers to obtain a better estimate of the
⫺0·5 hazard associated with aftershocks, and lead to new ap-
proaches to quantifying seismic performance and managing
Ground acceleration (AA02)
critical transport infrastructure, in which whole-life per-
khy (SS ⫹ RG) formance can be considered.
khy (SS)

Fig. 20. Comparison of predicted cumulative crest settlements CONCLUSIONS


(with and without re-grading) with centrifuge test measurements: In this paper, improved procedures for modelling the
test AA02 (Kobe) seismic response of dry granular shallow slopes using the
FEM and a fully strain-dependent Newmark sliding-block
procedure have been developed and validated against dy-
partially through undisturbed soil, thereby leading to an namic centrifuge test data. In the FE modelling, a non-
enhanced average frictional resistance along the slip plane. linear elasto-plastic constitutive model was used, the para-
meters of which could be estimated based only on relative
density using existing correlations, or using routine labora-
COMPARISON OF SLIDING-BLOCK AND FEM tory tests to develop a soil-specific model. These methods,
Figure 21 presents a summary of the predictive ability of as they do not require specialist testing, offer significant
the FEM and the sliding-block models for determining benefits for use in routine design, although the predictions
permanent settlements. These have been grouped into ‘ex- obtained are of an approximate nature. It was demonstrated
isting’ models, representing the current state-of-the-art, that the use of soil-specific parameters gave far improved
namely sliding block with strain-softening and FEM using predictions, particularly of permanent settlement, compared
the previously published parameters of Brinkgreve et al. with the existing correlations, which overpredicted settle-
(2010), and ‘improved’ models, consisting of the fully ments, particularly in subsequent earthquakes. This high-
lights the value of performing adequate site investigation.
In the improved sliding-block model, reduction in slope
12
Improved methods:
angle with slip/strain (re-grading or geometric hardening)
Sliding block (SS ⫹ RG)
has been incorporated alongside an existing strain-softening/
Predicted/measured (crest settlement)

10 hardening formulation.
FEM (ii)
The sliding-block and FE models gave comparable predic-
Existing methods: tions of permanent slip, capturing the significant decay in
8
Sliding block (SS) ground displacement (geometric hardening) with subsequent
FEM (i) shaking observed in the centrifuge tests. They therefore
6
Parity
permit both the response of slopes under strong aftershocks,
and the whole-life performance of a slope, to be quantified.
Whereas the sliding-block models are useful in preliminary
4
design, owing to the limited soil property data required and
the reduced computational effort, the FEM is additionally
2 able to quantify the dynamic performance of the soil and the
ground deformation profile (angular distortion) at the crest.
This would provide the information necessary to make a
0 detailed study of the seismic hazard posed to infrastructure
1 2 3 4
located at the slope crest, without requiring an excessive
Earthquake no.
amount of specialist laboratory testing, and is therefore
Fig. 21. Accuracy of ‘existing’ models/procedures, compared with complementary to the sliding-block models, being useful in
those proposed in this and the companion paper (‘improved’ the later stages of detailed design. The FE models generally
models) for predicting permanent crest settlement overpredicted the magnitude of dynamic ground motions in
AFTERSHOCKS AND THE WHOLE-LIFE SEISMIC PERFORMANCE OF GRANULAR SLOPES 1243
the slopes (particularly for components above 3 Hz). Adding 9pk (secant) peak angle of friction
in additional damping using the Rayleigh formulation im- ł9 effective angle of dilation
proved this, but had an adverse effect on the prediction of
permanent slope movements.

REFERENCES
ACKNOWLEDGEMENTS Ambrayses, N. & Srbulov, M. (1995). Earthquake induced displace-
The authors would like to express their sincere gratitude ments of slopes. Soil Dynam. Earthquake Engng 14, No. 1,
to Mark Truswell and Colin Stark at the University of 59–71.
Dundee for their assistance in performing the centrifuge Anastasopoulos, I., Gazetas, G., Bransby, M., Davies, M. & El
tests. The first author would also like to acknowledge the Nahas, A. (2007). Fault rupture propagation through sand: finite-
financial support of his PhD studies from the Ministry of element analysis and validation through centrifuge experiments.
Higher Education and Scientific Research (MOHESR) of the J. Geotech. Geoenviron. Engng 133, No. 8, 943–958.
Republic of Iraq. Ashford, S. A., Sitar, N., Lysmer, J. & Deng, N. (1997). Topo-
graphic effects on the seismic response of steep slopes. Bull.
Seismol. Soc. Am. 87, No. 3, 701–709.
Benz, T. (2006). Small-strain stiffness of soils and its numerical
NOTATION consequences. PhD thesis, University of Stuttgart, Germany.
a ground acceleration Bertalot, D. (2012). Behaviour of shallow foundations on layered
ag underlying bedrock peak acceleration soil deposits containing loose saturated sands during earth-
aslip slip acceleration quakes. PhD thesis, University of Dundee, UK.
Cu coefficient of uniformity Bertalot, D. & Brennan, A. J. (2012). Influence of bearing pressure
C2 coefficient of curvature on liquefaction-induced settlement of shallow foundations.
c9 cohesion intercept Géotechnique 63, No. 5, 391–399, http://dx.doi.org/10.1680/
ck stiffness-proportional Rayleigh damping coefficient geot.11.P.040.
cm mass-proportional Rayleigh damping coefficient Bertalot, D., Brennan, A. J., Knappett, J. A., Muir Wood, D. &
D10 particle diameter at which 10% is smaller Villalobos, F. A. (2012). Use of centrifuge modelling to improve
D30 particle diameter at which 30% is smaller lessons learned from earthquake case histories. Proceedings of
D60 particle diameter at which 60% is smaller the 2nd European conference on physical modelling in geotech-
d slope parallel slip nics, Eurofuge 2012, Delft, the Netherlands.
Eoed oedometric tangent stiffness (in compression) Bolton, M. D. (1986). The strength and dilatancy of sands.
Eur unloading–reloading stiffness Géotechnique 36, No. 1, 65–78, http://dx.doi.org/10.1680/geot.
E50 triaxial secant stiffness (at 50% of deviatoric failure stress 1986.36.1.65.
in drained triaxial compression) Bouckovalas, G. D. & Papadimitriou, A. G. (2005). Numerical
e natural void ratio evaluation of slope topography effects on seismic ground motion.
emax maximum void ratio Soil Dynam. Earthquake Engng 25, No. 7–10, 547–558.
emin minimum void ratio Bransby, M. F., Brown, M. J., Knappett, J. A., Hudacsek, P.,
F static factor of safety Morgan, N., Cathie, D., Maconochie, A., Yun, G., Ripley, A. G.,
fn natural frequency Brown, N. & Egborge, R. (2011). Vertical capacity of grillage
G shear modulus foundations in sand. Can. Geotech. J. 48, No. 8, 1246–1265.
G0 small-strain modulus Brennan, A. J. & Madabhushi, S. P. G. (2009). Amplification of
Gs specific gravity seismic accelerations at slope crests. Can. Geotech. J. 46, No. 5,
g acceleration due to gravity (¼ 9.81 m/s2 ) 585–594.
H slope height above toe Brennan, A. J., Thusyanthan, N. I. & Madabhushi, S. P. G. (2005).
ID relative density Evaluation of shear modulus and damping in dynamic centrifuge
K0 lateral earth pressure coefficient (at rest) tests. J. Geotech. Geoenviron. Engng 131, No. 12, 1488–1497.
kh pseudo-static seismic horizontal acceleration Brinkgreve, R. B. J., Engin, E. & Engin, H. K. (2010). Validation
khy yield acceleration of empirical formulas to derive model parameters for sands. In
Mw moment magnitude Numerical methods in geotechnical engineering (eds T. Benz
m power-law index for stress-level dependence of stiffness and S. Nordal), pp. 137–142. Rotterdam, the Netherlands: CRC
PGA peak ground acceleration (at soil surface) Press/Balkema.
pref reference stress (¼ 100 kPa) BSI (1990). BS 1377, Part 4: Maximum and minimum dry density
Rf ratio of deviatoric failure stress to asymptotic limiting for granular soil. London, UK: British Standards Institution.
deviator stress BSI (2005a). EN 1998-1: 2004: Eurocode 8: Design of structures
S soil factor (Eurocode 8) for earthquake resistance – Part 1: General rules, seismic
ST topographic amplification factor (Eurocode 8) actions and rules for buildings. London, UK: British Standards
t shear band thickness Institution.
u pore water pressure BSI (2005b). EN 1998-5: 2004: Eurocode 8: Design of structures
Vs shear wave velocity for earthquake resistance – Part 5: Foundations, retaining struc-
z depth of slip plane tures and geotechnical aspects. London, UK: British Standards
 slope angle Institution.
0 initial slope angle (pre-earthquake) Hardin, B. O. & Drnevich, V. P. (1972). Shear modulus and
ª soil unit weight damping in soils: design equations and curves. J. Soil Mech.
ªd dry unit weight Found. Div., ASCE 98, No. SM7, 667–692.
ªsat saturated unit weight Ishibashi, I. & Zhang, X. (1993). Unified dynamic shear moduli
s shear strain and damping ratios of sand and clay. Soils Found. 33, No. 1,
s,cs shear strain at critical state 182–191.
s,pk shear strain at peak state Jibson, R. W. (1993). Predicting earthquake-induced landslide dis-

add viscous damping ratio placements using Newmark’s sliding block analysis. Transp. Res.
ur Poisson’s ratio (unload–reload) Record 1411, 9–17.
9 normal effective stress Khazai, B. & Sitar, N. (2004). Evaluation of factors controlling
applied applied shear stress earthquake-induced landslides caused by the Chi-Chi earthquake
ult shear strength and comparison with the Northridge and Loma Prieta events. J.
9 effective angle of friction Engng Geol. 71, No. 1–2, 79–95.
9cs critical state angle of friction Kim, J. & Sitar, N. (2004). Direct estimation of yield acceleration
1244 AL-DEFAE, CAUCIS AND KNAPPETT
in slope stability analyses. J. Geotech. Geoenviron. Engng 130, Newmark, N. M. (1965). Effects of earthquakes on dams and
No. 1, 111–115. embankments. Géotechnique 15, No. 2, 139–159, http://dx.doi.
Knappett, J. A. & Craig, R. F. (2012). Craig’s soil mechanics, 8th org/10.1680/geot.1965.15.2.139.
edn. London, UK: Spon Press. Oda, M. & Kazama, H. (1998). Microstructure of shear bands and
Kramer, S. L. (2008). Performance-based earthquake engineering: its relation to the mechanism of dilatancy and failure of dense
opportunities and implications for geotechnical engineering prac- granular soils. Géotechnique 48, No. 4, 465–482, http://dx.doi.
tice. Keynote paper. Proceedings of the 4th conference on org/10.1680/geot.1998.48.4.465.
geotechnical earthquake engineering and soil dynamics, pp. 1– Rathje, E. M. & Bray, J. D. (1999). An examination of simplified
32. Reston, VA, USA: ASCE. earthquake-induced displacement procedures for earth structures.
Lauder, K. (2011). The performance of pipeline ploughs. PhD Can. Geotech. J. 36, No. 1, 72–87.
thesis, University of Dundee, UK. Rowe, P. W. (1962). The stress–dilatancy relation for static equili-
Lo Presti, D. C. F., Pallara, O., Fioravante, V. & Jamiolkowski, M. brium of an assembly of particles in contact. Proc. R. Soc. A
(1998). Assessment of quasi-linear models for sands. In Pre- 269, No. 1339, 500–527.
failure deformation behaviour of geomaterials (eds R. Jardine, Santos, J. A. & Correia, A. G. (2001). Reference threshold shear
M. Davies, D. Hight, A. Smith and S. Stallebrass), pp. 363–372. strain of soil: its application to obtain a unique strain-dependent
London, UK: Thomas Telford. shear modulus curve for soil. Proceedings of the 15th interna-
Loukidis, D., Bandini, P. & Salgado, R. (2003). Stability of tional conference on soil mechanics and geotechnical engineer-
seismically loaded slopes using limit analysis. Géotechnique 53, ing, Istanbul, Turkey, vol. 1, pp. 267–270.
No. 5, 463–479, http://dx.doi.org/10.1680/geot.2003.53.5.463. Schanz, T., Vermeer, P.A. & Bonnier, P.G. (1999). The hardening-
Lysmer, J. & Kuhlmeyer, R. L. (1969). Finite dynamic model for soil model: formulation and verification. In Beyond 2000 in
infinite media. Proc. ASCE, J. Engng Mech. Div. 95, No. 4, computation geotechnics (ed. R. B. J. Brinkgreve), pp. 281–290.
859–887. Rotterdam, the Netherlands: Balkema.
Matasovic, N., Kavazanjian, E. Jr & Yan, L. (1997). Newmark Smith, C. C. & Gilbert, M. (2007). Application of discontinuity
deformation analysis degrading yield acceleration. Proceedings layout optimization to plane plasticity problems. Proc. R. Soc. A
of Geosynthetics ’97, Long Beach, CA, USA, vol. 2, pp. 989– 463, No. 2086, 2461–2484.
1000. Wartman, J., Bray, J. & Seed, R. (2003). Inclined plane studies of
Muir Wood, D. (2002). Some observations of volumetric instabil- the Newmark sliding block procedure. J. Geotech. Geoenviron.
ities in soils. Int. J. Solids Struct. 39, No. 13–14, 3429–3449. Engng 129, No. 8, 673–684.
Mülhaus, H. B. & Vardoulakis, I. (1987). The thickness of shear Wartman, J., Seed, R. & Bray, J. (2005). Shaking table modeling of
bands in granular soils. Géotechnique 37, No. 3, 271–283, seismically induced deformations in slopes. J. Geotech. Geoen-
http://dx.doi.org/10.1680/geot.1987.37.3.271. viron. Engng 131, No. 5, 610–622.

Вам также может понравиться