Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

EARTHQUAKE ENGINEERING AND STRUCTURAL DYNAMICS

Earthquake Engng Struct. Dyn. 2009; 38:679–698


Published online 9 March 2009 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/eqe.902

Performance evaluation of a nonlinear Winkler-based shallow


foundation model using centrifuge test results

Prishati Raychowdhury1, ‡ and Tara C. Hutchinson2, ∗, †, §


1 Earth Mechanics, Inc., 17660 Newhope Street, Suite E, Fountain Valley, CA-92708, U.S.A.
2 Department of Structural Engineering, University of California, San Diego, La Jolla, CA-92093, U.S.A.

SUMMARY
When a structure supported on shallow foundations is subjected to inertial loading due to earthquake
motions, the foundation may undergo sliding, settling and rocking movements. Even if the capacity of
the foundation is mobilized, the soil–foundation interface may dissipate significant amounts of vibrational
energy, resulting in reduced force demands to the superstructure. If the capacity is not mobilized, these
movements introduce additional flexibility to the system, which may shift its period away from the
potentially hazardous zone of response spectra for most earthquake ground motions. In either case, transient
and permanent deformations will need to be accounted for.
To consider the aforementioned benefits and consequences in performance-based seismic design, robust
and reliable numerical modeling tools are needed. In this article, a Winkler-based modeling framework is
proposed to address this issue. The model includes a distributed array of mechanistic nonlinear inelastic
springs, dashpots, and gap elements, with backbone curves of the nonlinear springs calibrated against
shallow foundation experiments. Model evaluation is conducted by simulating the response of a number
of centrifuge experiments. Experiments considered include square and strip footings, bridge and building
models, static and dynamic loading, footings on sand and clay, a range of static vertical factors of safety,
and a range of aspect ratios. It is observed that the model can reasonably predict measured footing response
in terms of moment, shear, settlement and rotational demands. In addition, the general hysteresis shape
of the moment–rotation, settlement–rotation and shear–sliding curves is reasonably captured. Copyright
q 2009 John Wiley & Sons, Ltd.

Received 17 June 2008; Revised 1 November 2008; Accepted 5 January 2009

KEY WORDS: shallow foundation; centrifuge tests; BNWF modeling; nonlinear analysis

∗ Correspondence to: Tara C. Hutchinson, Department of Structural Engineering, University of California, San Diego,
La Jolla, CA-92093, U.S.A.

E-mail: tara@ucsd.edu

Staff Engineer.
§ Associate Professor.

Contract/grant sponsor: NSF (PEER); contract/grant number: EEC-9701568

Copyright q 2009 John Wiley & Sons, Ltd.


680 P. RAYCHOWDHURY AND T. C. HUTCHINSON

1. INTRODUCTION

Nonlinear soil–structure interaction (SSI) of shallow foundations subjected to earthquake ground


motions may include settlement, sliding, and rocking of the foundation, gap formation between the
footing and soil, and energy dissipation from hysteretic effects. Capturing these behaviors remains
a key challenge to the earthquake engineering community, particularly in light of the desire to
embrace performance-based design. Of particular interest are the energy dissipative and re-centering
benefits of a rocking shallow foundation (e.g. early work of Housner [1]). Recent experimental
evidence has further sparked the interest of the community in utilizing these benefits (e.g. [2–13]).
However, the consequences of allowing a shallow foundation to yield and/or uplift, namely,
transient and permanent settlement, sliding and rotation, must be reasonably estimated in design.
Design and rehabilitation provisions have been available for many years to characterize SSI
effects when conducting structural analyses (e.g. [14–18]). However, these design documents
have largely focused on supporting simplified pseudo-static force-based or pushover-type proce-
dures, with the soil–foundation interface characterized in terms of elastic impedance functions that
prescribe stiffness and damping characteristics. For example, ATC–40 [14] accounts for SSI by
using Winkler springs with elastic stiffness suggested by Gazetas (e.g. [19, 20]), whereas NEHRP–
2000 [15] and ASCE–7 05 [18] account for SSI using an increased period and modified damping
ratio of the soil–structure system. Such approaches are not able to capture the aforementioned
complex behaviors.
Over the years, numerous studies have been conducted to model the behavior of structures
supported on shallow foundations (e.g. [21–32]). A number of these have focused on Winkler-
based approaches, whereby a system of closely spaced independent springs are placed along
the footing length and attached to structural footing elements. The Winkler spring approach is
appealing to engineering practitioners due to its simplicity, minimal computational effort and
ease of implementation. In addition, the mechanistic nature of spring-based models can simplify
calibration, if isolated response data exist. However, it is now widely accepted that any calibrated
mechanistic spring behavior must be able to capture salient physical behavior of the foundation
such as uplift and permanent settlement.
A notable early study incorporating the use of nonlinear elastic–plastic Winkler springs to model
shallow foundations was conducted by Chopra and Yim [22, 23]. A few years later, Nakaki and
Hart [24] used elastic, no-tension Winkler springs with viscous dampers to model the response
of an inelastic shearwall. Nakaki and Hart report that ductility demands are significantly reduced
if uplifting at the base of the foundation is accounted for. In an application to bridge modeling,
Fenves [26] developed a composite spring model that was well suited for capturing the cumulative
permanent settlement observed due to rocking of shallow footings under large amplitude cyclic
motions. Harden et al. [29] developed a Winkler-based model using pile-calibrated nonlinear
backbone curves to model the behavior of shallow strip footings supporting rocking-dominated
shearwall buildings. Recently, Allotey and Naggar [28, 32] developed a Winkler-based approach
utilizing multi-linear, no-tension backbone curves. The authors compare the model with centrifuge
tests by Gajan et al. [8].
In this paper, a beam-on-nonlinear-Winkler-foundation (BNWF) approach for modeling shallow
foundation behavior under seismic loading is proposed. The model consists of a system of closely
spaced independent, nonlinear springs, coupled with a dashpot and gap elements, with backbone
curves of the nonlinear springs calibrated against shallow foundation experiments. Vertical springs
distributed along the base of the footing are intended to capture the rocking, uplift and settlement,

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
PERFORMANCE EVALUATION OF A SHALLOW FOUNDATION MODEL 681

whereas horizontal springs attached to the sides of the footing are intended to capture the resistance
to sliding and passive pressure. The proposed model builds upon that of Harden et al. [29], and
complements that of Allotey and Naggar [28, 32]. The model has the following attributes: (i)
backbone curves of the spring models are calibrated against shallow footing tests, (ii) elastic
stiffnesses after [19] are directly adopted to characterize the spring curve, (iii) nominal tension
capacity is allowed in the model, (iv) a variable vertical stiffness distribution along the base of
the model is incorporated to capture rotational stiffness, and (v) the model is able to capture
experimentally observed behavior for a broad range of shallow footing, soil types, vertical factors
of safety and loading histories.

2. METHODOLOGY

2.1. Model framework


The proposed two-dimensional BNWF model is constructed with a mesh of closely spaced, inde-
pendent nonlinear spring elements placed vertically along the length of the footing and horizontally
at the ends of the footing (Figures 1(a) and (b)). The proposed model is an extension of earlier
studies described in [29, 33]. The array of vertical q–z springs is intended to capture vertical and
rotational resistance of the footing, whereas two springs, namely p–x and t–x, are placed horizon-
tally to capture passive and sliding resistance of the footing, respectively. A region of increased
stiffness is provided at the end of the footing (Figure 1(c)), such that the rotational stiffness is
appropriately accounted for [29, 33]. The model has been developed within the OpenSees platform
[34] and implemented as a mesh generator. Individual q–z, p–x, and t–x mechanistic springs are
modeled with one-dimensional zero-length elements and are attached to beam–column elements,
which represents the stiffness of the structural footing. As desired, an analyst may connect to any
point along the footing with representative superstructure elements.

2.2. Mechanistic springs


The constitutive relations used for the q–z, p–x and t–x springs are represented by nonlinear back-
bone curves that have been calibrated against shallow foundation tests [35]. Within the OpenSees
framework, these are the QzSimple2, PxSimple1 and TxSimple1 material models, respectively. The
models are variations of earlier, pile-calibrated backbone curves developed by Boulanger [36] based
on early material model development described in Boulanger et al. [37]. The QzSimple2, PxSimple1
and TxSimple1 material models differ from their parent (pile-calibrated) models (QzSimple1,
PySimple1 and TzSimple1) only in the backbone shape parameters.
Each material captures the ‘far-field’ elastic behavior and ‘near-field’ permanent displacements.
The material models are mechanistic, based on an arrangement of various linear and nonlinear
springs, gap elements and a dashpot, as shown in Figure 2(a). Radiation damping can be accounted
for via a dashpot that is placed in parallel with the far-field elastic component. The backbone
curves are thus characterized by a linear elastic region, then an increasingly growing nonlinear
region. The QzSimple2 material has an asymmetric hysteretic response, with a backbone curve
defined by an ultimate load on the compression side and a reduced strength in tension to account
for soil’s low strength in tension. The PxSimple1 is envisioned to capture the passive resistance,
associated stiffness and potential gapping of embedded shallow footings subjected to lateral loads.
This material model is characterized by a pinched hysteretic behavior, which more suitably can

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
682 P. RAYCHOWDHURY AND T. C. HUTCHINSON

Figure 1. Proposed BNWF model: (a) hypothesized footing–structural system; (b) idealized model; and
(c) variable vertical stiffness distribution (after [29, 33]).

Elastic
Region Plastic
Region
1.2

0.8
q/qult

q50/qult
0.4
Cr

0
0 z0 z50 4 8 12
(a) (b) z/z50

Figure 2. (a) Conceptual construction of the mechanistic springs (after Boulanger et al. [37]) and
(b) backbone curve for q–z material model.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
PERFORMANCE EVALUATION OF A SHALLOW FOUNDATION MODEL 683

Figure 3. Cyclic response of: (a) q–z spring (Qzsimple2 material model); (b) p–x spring (Pxsimple1
material model); and (c) t–x spring (Txsimple1 material model).

account for the phenomena of gapping during unloading on the opposite side of the footing. The
TxSimple1 material is intended to capture the frictional resistance along the base of a shallow
foundation. This material is characterized by a large initial stiffness and a broad hysteresis, as
anticipated for frictional behavior associated with foundation sliding. The cyclic response of the
material models, when subjected to a sinusoidal displacement is demonstrated in Figure 3. This
figure demonstrates the progression of the monotonic curve to a cyclic load-deformation response,
and in particular shows the gap growth capabilities of the models. Furthermore, this figure shows the
reduced strength in tension and corresponding asymmetric behavior of QzSimple2, the symmetric
behavior of PxSimple1 and TxSimple1 and the broad hysteresis of the sliding resistance (without
gapping) provided by TxSimple1.
The equations used to describe the QzSimple1 material model are similar to those used for the
PySimple1 materials described in [37]. These same equations are used for the material models
herein, with the exception of modified shape parameters. For completeness of presentation, they
are described briefly.
In the linear elastic region, the instantaneous load q is assumed to be linearly proportional with
the instantaneous displacement z via the initial elastic (tangent) stiffness kin , i.e.

q = kin z (1)

The range of the elastic region is defined by the following relation:

qo = Cr qult (2)

where qo = load at the yield point and Cr = parameter controlling the range of the elastic portion.
In the nonlinear (post-yield) portion, the backbone curve is described by
 n
cz 50
q = qult −(qult −qo ) p (3)
cz 50 +|z p − z o |
where qult = ultimate load, z 50 = displacement at which 50% of the ultimate load is mobilized,
p
qo = load at the yield point, z o = displacement at the yield point, z p = displacement at any point in
the post-yield region and c and n are the constitutive parameters controlling the shape of the post-
yield portion of the backbone curve. The gap component of the spring is a parallel combination
of a closure and drag spring (Figure 2(a)). The closure component (q c − z g ) is a bilinear elastic

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
684 P. RAYCHOWDHURY AND T. C. HUTCHINSON
Normalized vertical load (Q/Qult)

Normalized lateral load (V/Pult)

Normalized lateral load (V/Tult)


1.2 1.2 1.2

0.8 0.8 0.8

Shallow footing calibrated ( μ) Shallow footing calibrated ( μ)


Shallow footing calibrated ( μ)
0.4 0.4 0.4 Shallow footing calibrated ( μ±σ)
Shallow footing calibrated ( μ±σ) Shallow footing calibrated ( μ±σ)
Pile calibrated (Boulanger [36]) Pile calibrated (Boulanger [36]) Pile calibrated (Boulanger [36])

0 0
0 2 4 6 8 10 0
0 2 4 6 8 0 2 4 6 8 10
(a) Normalized settlement (s/z50) (b) Normalized lateral displacement (u/x50) (c) Normalized lateral displacement (u/x50)

Figure 4. Regressed backbones for the (a) q–z; (b) p–x; and (c) t–x material models.

spring, which is relatively rigid in compression and very flexible in tension. The nonlinear drag
component (q d − z g ) of the curve is controlled by the following equation:
 
z 50
qd = Cd qult −(Cd qult −qo )
d
g (4)
z 50 +2|z g − z o |
where q d = drag force on the closure component, qod = q d at the start of the current loading cycle,
g
z o = z g at the start of the current loading cycle and Cd = ratio of the maximum drag (suction)
force to the ultimate resistance of the q–z material.
The expressions governing both PxSimple1 and TxSimple1 are quite similar to Equations (1)–(4),
with variations in the constants n, c and Cr , which control the general shape of the curve. Moreover,
both PxSimple1 and TxSimple1 are symmetric in compression and tension, as they represent the
cumulative behavior of the front and backside of the footing simultaneously, through a single
spring. Note that the same equations are valid for initial loading, unloading and reloading.
For calibration of QzSimple2, PxSimple1 and TxSimple1 material models, axial, sliding and
embedded laterally loaded footing tests [2, 4–6, 38, 39] are considered. By analyzing representative
models of the individual experiments, and iterating to obtain a best-fit to the experimental response,
the parameters describing the shape of the backbone curves are determined. The value of each
parameter that corresponds to the minimum residual normalized load (normalized with respect to
the ultimate load) is taken as the best-fit value. Once a set of best-fit values for each parameter and
each test is determined, statistics of that parameter are calculated (mean  and standard deviation
). Using the calculated  and ± values of each parameter (c, n and Cr ), a set of three nonlinear
backbone curves may be defined. Along with the mean backbone, a dispersion of behavior in terms
of mean ± standard deviation curves is also presented in Figure 4(a)–(c). The shallow foundation-
calibrated backbone curves used in this study are also compared with their parent (pile-calibrated)
backbone curves. Additional details of the calibration study are provided in [35, 40].

3. MODEL PERFORMANCE EVALUATION

To evaluate the BNWF model performance, numerical models are constructed of footing specimens
tested under 20-g centrifugal acceleration. The numerical models are then subjected to the same
input motions applied to the experimental specimens and comparison between the numerical and
experimental model results is undertaken. In the following, prototype units are used throughout,
unless otherwise noted.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
PERFORMANCE EVALUATION OF A SHALLOW FOUNDATION MODEL 685

3.1. Experiments considered in this study


Two sets of centrifuge test data have been selected for model evaluation. The first set of data is
from a series of building footings (primarily strip footings) supporting shearwalls. The second set
of data consists of isolated square footings supporting bridge columns. Experiments selected for
evaluation include both static cyclic tests, where inertial loads are applied at the superstructure, as
well as dynamically loaded model tests, where the base of the centrifuge model is shaken using
recorded earthquake motions or a cosine tapered motion. All experiments are carried out at the
University of California, Davis in the large 9 m radius NEES geotechnical centrifuge. The bridge
and building models considered for exercising the model are different from each other in the
following respects:

• All shearwall building models rest on strip footings with a B/L (width/length) ratio ≈ 0.2,
whereas the bridge model rests on two square footings, one with a width of 5.4 m and the other
with a width of 7.1 m.
• The range of vertical factor of safety, FSv , for the building footings considered is 2–14, whereas
the range of FSv for the bridge footings considered is 17–31.
• The input motions used for the building models are slow cyclic lateral pushes and cosine tapered
dynamic shakes, whereas the bridge footings were subjected to recorded earthquake ground
motions.
• Shearwall-footing models are resting on both sands and clays, whereas all bridge-footing models
are resting only on sands.
• The structural configuration for the building and bridge models are also different. The super-
structure of the shearwall models consists of a wall with uniform cross-sectional properties
supported by a footing, whereas the bridge column is a ‘lollipop’ structure with a deck mass
and column connected to a shallow square footing.

3.2. Shearwall-footing experiments


3.2.1. Experiment details. Table I presents the experimental details of shearwall-footing tests
considered for this study. Additional details may be found in [4–6, 8–10]. Some of the tests were
conducted on uniform fine-grained dense dry Nevada sand with a relative density Dr = 80% and
mean grain size D50 = 0.17 mm, whereas others were conducted on remolded San Francisco Bay
Mud (plastic limit, PL = 35 ∼ 40 and liquid limit, LL 88 ∼ 93) with an undrained shear strength cu of
clay 100±10 kPa. All tests considered are conducted with a strip footing of size 2.8 m×0.65 m, and
all but two of the experiments were conducted on surface-mounted footings. Embedded footings
were placed initially at 1B below the ground surface (for tests SSG03–03 and SSG03–07). The
static vertical factor of safety FSv for these tests ranged from 2.3 to 14.0 and the non-dimensional
moment to shear ratios M/H L ranged from 0.41 to 1.80.

3.2.2. Numerical modeling. Figure 5(a) shows a schematic diagram of the shearwall-footing exper-
imental specimens, including the overall geometry, and configuration of the various transducers.
Figure 5(b) shows the idealized model of the system. Note that the diagram is only shown schemat-
ically. In total 61 vertical q–z springs are placed along the length of the footing with a spacing of
2% of the total length of the footing. Two lateral springs, t–x and p–x, are used to provide sliding
and passive resistance, respectively. The vertical capacity of the footing is calculated using equa-
tions by Terzaghi [41], with depth, shape and inclination factors of Meyerhof [42]. The capacity

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
686 P. RAYCHOWDHURY AND T. C. HUTCHINSON

Table I. Details of the shearwall-footing tests considered for the BNWF model evaluation.
Test name Loading type Reference Mass (Mg) Soil type Df /B FSv M/H L
SSG04-06 Static [10] 68 Sand, Dr = 80% 0 2.3 1.20
SSG03-02 Static [9] 58 Sand, Dr = 80% 0 2.5 0.45
SSG02-05 Static [8] 58 Sand, Dr = 80% 0 2.6 1.72
SSG02-02 Static [8] 28 Sand, Dr = 80% 0 5.2 0.41
SSG02-03 Static [8] 28 Sand, Dr = 80% 0 5.2 1.75
SSG03-03 Static [9] 28 Sand, Dr = 80% 1 14.0 1.77
KRR03-03 Static [6] 36 Clay, cu = 100 kPa 0 2.8 1.80
SSG04-10 Dynamic [10] 36 Sand, Dr = 80% 0 4.0 1.80
SSG03-07 Dynamic [9] 58 Sand, Dr = 80% 1 7.2 1.80
KRR03-03 Dynamic [6] 36 Clay, cu = 100 kPa 0 2.8 1.70

AV 2 AV 1
Elastic beam
AH3 column
elements Mass

AH2 Lateral cyclic


CG loading (for
hs static tests) Zero-length
LH2
element
(t-x spring)
LV1 LV2 Nodes
hcg
LH1 G.S.
AH1
O
Df L
Zero-length Zero-length element
LH1, LH2 = Horizontal displacement transducers element (p-x spring for
LV1, LV2 = Vertical displacement transducers (q-z spring) Dynamic shaking embedded footing)
AV1, Av2 = Vertical accelerometer (for dynamic tests)
AH1, AH2, Ah3 = Horizontal accelerometer
(a) (b)

Figure 5. (a) Shearwall-footing test schematic diagram (after [30]) and (b) idealized model.

is distributed to individual vertical q–z spring based on its tributary area. The capacity on the
tension side is assumed to be 10% of the ultimate capacity of the compression side. The passive
resistance (for p–x spring) and sliding resistance (for t–x spring) are calculated after [43] and
assuming classical Mohr–Coulomb failure criteria, respectively. Vertical and lateral stiffnesses of
the springs are calculated using Gazeta’s equations [19]. It is important to note that certainly more
recent methods for calculating capacity are warranted for use as input into the model; however,
we recognize that the classical theories of, for example, Terzaghi in estimating bearing capacity
are used in nearly all situations in practice today. Nonetheless, the model is flexible enough to
accept as input, parameters such as bearing and lateral capacity, and/or stiffness, derived by other

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
PERFORMANCE EVALUATION OF A SHALLOW FOUNDATION MODEL 687

models (or determined via experiments), in the event alternative methods or data are desired.
A variable distribution of stiffness along the length of the footing is considered to account for
appropriate rotational stiffness. The stiffness intensity at the end region is increased by 2.5 times
that of the mid region. The length of the end region (increased stiffness region) is assumed to be
10% of the length of the footing. The specification of increased stiffness intensity zone is selected
based on recommendations by [29] for shallow strip footings with B/L ≈ 0.2. The shearwall and
the structural foundation are considered elastic and modeled as elastic beam–column elements.
Additional details of the input parameters and their selection protocol are provided in [35].

3.2.3. Comparison with experimental results. Figures 6 and 7 show comparison results for test
SSG04–06. This is a static cyclic test on a medium aspect ratio shearwall building (M/H L = 1.2)
with a vertical factor of safety, FSv = 2.3, resting on dense sand with Dr = 80%. The time history
responses of footing forces and displacements along with the input displacement for this test case
are shown in Figure 6, with the simulation shown in black and the experiment shown in gray. Note
that the time axis represents pseudo time as the test input is slow cyclic displacement at the top of
the structure. It can be observed that the experimentally observed behavior is well synchronized
and in good agreement with the simulation results in terms of capturing salient features such
as maximum moment, shear, rotation and settlement demands. However, sliding displacement
demands are under-predicted, due to the over-prediction of shear capacity, which as a result is not
mobilized during the simulation. An increase in moment demand at the footing should result in a
decrease in shear capacity; however, this cannot be captured if the lateral and vertical springs are
uncoupled.
Figure 7(a)–(d) show moment versus rotation, settlement versus rotation, shear force versus
sliding and settlement versus sliding histories, respectively. It is observed that the model compares

Figure 6. Time history responses for test SSG04–06: (a) moment; (b) shear; (c) sliding; (d) settlement;
(e) rotation; and ( f ) input displacement.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
688 P. RAYCHOWDHURY AND T. C. HUTCHINSON

Figure 7. Footing response comparison for test SSG04–06 (static cyclic test on dense sand
with Dr = 80%, FSv = 2.3 and M/H L = 1.20): (a) moment–rotation; (b) settlement–rotation;
(c) shear–sliding; and (d) settlement–sliding.

reasonably well with the experimental results in terms of capturing the shapes of the hysteresis
loops and rotational and lateral stiffnesses. Additional results showing the time histories of footing
responses are provided in [35].
Table II summarizes the absolute maximum force and displacement demands observed in the
experiment and simulated using the BNWF model for the shearwall-footing test cases described
in Table I. Force demands of interest include the absolute maximum moment |Mmax | and shear
|Vmax |, whereas the displacement demands of interest include the absolute maximum rotation
|max |, sliding |u max | and settlement |smax |. The deviation in the experimentally observed and
the simulated demands is also summarized in the same table (last five columns) along with the
mean deviation predicting each demand parameter. The parameters |ε M |, |ε |, |εV |, |εu | and |εs |
represent the absolute deviation of the simulated and the experimental peak demands normalized
by the experimental peak demands for each of the moment, rotation, shear, sliding and settlement
responses, respectively. The mean deviations of the numerical simulations, in terms of peak moment,
shear, settlement and rotation from the experimental results are: 9, 16, 14 and 11%, respectively,
whereas the mean deviation in estimating the sliding demand is 44%.
Figure 8 shows the summary of predictive capability of BNWF model, indicating the over- or
under-prediction of experimentally observed values of each demand parameter. It is noted that
the moment, shear and rotation do not follow any particular trend; in some of the cases they
are under-predicted, whereas in other cases, they are slightly over-predicted; however, the mean

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
Copyright q
Table II. Comparison of experimental and simulated demands for shearwall-footing tests.
Experimental values BNWF simulation Absolute percent deviation
|Mmax | |max | |Vmax | |u max | |smax | |Mmax | |max | |Vmax | |u max | |smax | | M | | | |V | |u | |s |

2009 John Wiley & Sons, Ltd.


Test name (kN m) (rad) (MN) (mm) (mm) (kN m) (rad) (MN) (mm) (mm) (%) (%) (%) (%) (%)
SSG04-06 560 0.050 154 54 307 564 0.047 168 30.8 277 1 6 9 43 10
SSG03-02 306 0.014 226 52 237 224 0.01 182 35 173 27 29 19 33 27
SSG02-05 530 0.033 108 23 137 511 0.035 106 23 120 4 6 2 0 12
KRR03-02 356 0.069 69 68 117 300 0.08 67 55 93 16 16 3 19 21
KRR03-03 287 0.012 No data No data 8 241 0.008 No data No data 8 16 33 — — 4
SSG04-10 278 0.009 54 8 45 273 0.009 30 2.2 47 2 0 44 73 4
SSG02-02 189 0.014 158 67 126 218 0.015 170 34 115 15 7 8 49 9
SSG02-03 383 0.063 71 48 83 355 0.065 75 37 91 7 3 6 23 10
SSG03-07 527 0.02 100 14 88 540 0.013 60 2.2 78 2 35 40 84 11
SSG03-03 436 0.063 83 21 36 428 0.063 91 5 34 2 0 10 76 6
Mean 9 14 16 44 11
PERFORMANCE EVALUATION OF A SHALLOW FOUNDATION MODEL

DOI: 10.1002/eqe
Earthquake Engng Struct. Dyn. 2009; 38:679–698
689
690 P. RAYCHOWDHURY AND T. C. HUTCHINSON

Figure 8. Summary of performance of the proposed BNWF model in


predicting shearwall-footing test results.

absolute deviation is less than 20%. Settlement demand is predicted very well for all of the cases
except for two cases where the maximum settlement is very large, exceeding 200 mm. Sliding on
the other hand is under-predicted in nearly all cases. This may be due to the lack of coupling
between the vertical and lateral modes of response inherent in the model.
An important set of parameters from a foundation design perspective is the residual deformations
at the end of an earthquake ground excitation, i.e. residual rotation res , sliding u res and settlement
sres . Table III summarizes a comparison of experimental and simulation values for these parameters
along with their deviations. The parameters  res , u res and s res represent the deviation of the
absolute experimental and simulated residual values (experimental minus simulated) for each of
the rotation, sliding and settlement responses, respectively. Residual rotation values are in general
very small. For sliding, similar to the peak demands, residuals are under-predicted in general and
also have associated large percentage deviations. Residual settlements on the other hand are very
well predicted as one can observe from the percentage deviation in predicting residual settlement
values (maximum value of 28.8%). One can expect that the maximum residual settlement occurs
near the end of the loading history and thus is either equal or very close to the peak settlement
(see Figure 5(d)).

3.3. Bridge-footing experiments


3.3.1. Experiment details. The bridge-footing experiments considered are those conducted by
Ugalde et al. [13] on embedded square footings resting on dense dry Nevada sand of relative density

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
PERFORMANCE EVALUATION OF A SHALLOW FOUNDATION MODEL 691

Table III. Comparison of experimental and simulated residual responses for shearwall-footing tests.
Experimental values BNWF simulation Deviation
|res | |u res | |sres | |res | |u res | |sres | res ures sres
Test name (rad) (mm) (mm) (rad) (mm) (mm) (rad) (mm) (mm)
SSG04-06 0.004 41.25 307.50 0.011 1.25 277.20 −0.007 40.00 30.30
SSG03-02 0.001 9.60 240.00 0.002 6.01 171.00 −0.001 3.59 69.00
SSG02-05 0.027 1.43 132.00 0.028 3.11 120.00 −0.001 −1.69 12.00
KRR03-02 0.000 23.11 111.60 0.002 32.68 92.73 −0.002 −9.57 18.87
KRR03-03 0.000 — 7.00 0.001 — 7.90 −0.001 — −0.90
SSG04-10 0.002 0.27 45.06 0.002 0.15 46.19 0.000 0.12 −1.13
SSG02-02 0.011 4.45 125.00 0.007 6.28 114.40 0.004 −1.83 10.60
SSG02-03 0.011 40.07 82.67 0.009 31.89 90.15 0.001 8.18 −7.48
SSG03-07 0.015 4.00 87.26 0.003 0.49 77.61 0.012 3.51 9.65
SSG03-03 0.003 18.48 36.10 0.001 3.65 34.09 0.002 14.83 2.01

Table IV. Details of the bridge-footing tests considered for the BNWF model evaluation.
Station name Test name Mass (Mg) L(m) B(m) Df /B FSv M/H L
E JAU01-E05 to E08 1100 5.4 5.4 0.3 17.0 2.50
F JAU01-F05 to F08 1172 7.1 7.1 0.24 31.0 1.90

80%. Specimens are scaled versions of typical bridge-footing configurations used by the California
Department of Transportation. The prototype structure is a typical reinforced concrete single
column bridge bent modeled as a ‘lollipop’ structure, with a deck mass and column connected to a
shallow spread footing. Footings are square with widths of three or four times the diameter of the
column (prototype column diameters = 1.8 m) and embedded either 0.24 or 0.30 times the breadth
of the footing. The static vertical factor of safety FSv for these models is much higher than that of
the shearwall-footing models, namely 17 and 31. These large FSv values are based on allowable
bearing pressures, however, these structures are designed to minimize settlements. The normalized
moment to shear ratios of the models considered are 1.90 and 2.50. Structures at Stations E and F
were subject to dynamic loading at the base of their containers, using the shaking table mounted on
the centrifuge. Ground motions imposed on the model container were scaled and filtered motions
from recordings in the Tabas 1978 earthquake and a Los Gatos recording of the 1989 Loma Prieta
earthquake. These motions come from near-field records generated for the SAC Steel Project (see
http://nisee.berkeley.edu/data/strong motion/sacsteel/ ground motions.html). Shaking events five,
six and eight were chosen for the model evaluation because they are of moderate intensity; events
1–4 and 7 resulted in very little settlement or nonlinear load-deformation of the footing. Table IV
presents the list of experiments with their details.

3.3.2. Numerical modeling. Figure 9(a) shows a schematic diagram of the bridge-footing exper-
imental specimens, including the overall geometry, and configuration of the various transducers.
Figure 9(b) shows the idealized model of the system using the proposed model. The diagram is
only shown schematically. Numerical model of shallow foundation in this case is similar to that

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
692 P. RAYCHOWDHURY AND T. C. HUTCHINSON

Mass

Elastic beam
column
elements

(a)
Zero-length
element
(t-x spring)
Nodes

Zero-length Zero-length
element element
(q-z spring) Dynamic shaking (p-x spring)
(b) (c)

Figure 9. (a) and (b) Bridge-footing test schematic diagram (after [13]); and (c) idealized model.

of shearwall-footing cases, except that the stiffness intensity at the end region is increased by five
times that of the mid region and the length of the end region is assumed to be 16% of the length
of the footing, based on recommendations of [29] for square footings. To model the structural
foundation and the bridge column, elastic beam–column elements are used. In these models, the
free-field motions recorded near the base level of the footings were used directly as input to the
base of the model in the simulation. Kinematic interaction between the footing and soil is ignored.

3.3.3. Comparison with experimental results. Figures 10 and 11 show comparison results from the
analysis of the bridge-footing tests for test cases JAU01–E08 and JAU01–F08. Figure 10, which
shows the time history response of the experiments and simulations, indicates that the BNWF
model is able to capture peak moment, rotation and settlement demands reasonably well with a
maximum absolute deviation of 14, 21 and 13%, respectively. Figure 11 shows the moment–rotation,
settlement–rotation and the input motion characteristics, for this analysis case. It is observed that
the overall prediction of footing demands is quite reasonable for both footings, in terms of the
shapes of the moment–rotation and settlement–rotation curves. It should be noted that in each
subsequent event, there is some initial settlement, which is accrued from the previous loading
cycles. The magnitude of this deformation was applied as an initial displacement, when simulating
a particular event.
Table V presents a summary of the peak footing force and displacement demands predicted by
the BNWF model compared with those measured during the experiments. One can observe that

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
PERFORMANCE EVALUATION OF A SHALLOW FOUNDATION MODEL 693

Figure 10. Time history responses for bridge-footing tests (tests JAU01–E08—left
column and JAU01–F08—right column).

the mean deviation of predicted peak moment, rotation and settlement demands is 21, 36 and 20%,
respectively.
It is observed from Figure 12(a) that the moment demand on the footing is under-predicted by
the BNWF model for both Stations F and E for all events. The range of deviation in predicting peak
moment demand is 13–25% with a mean of 21%. This may be due to the uncertainty in estimating
the bearing capacity of the soil. During testing, the soil underneath the footing is compacted
and densified during previous loading events and therefore has a greater capacity than estimated
theoretically. Another reason for the under-estimation of moment capacity may be due to ignoring
the side friction contribution to moment capacity of the embedded footing. This additional capacity
would be a larger contributor to the overall footing capacity for square footings. Note also that the
footing settles during each event, becoming embedded deeper and deeper, therefore increasing the
overall capacity as well. As observed from Figure 12(b), the rotation is under-estimated between
21 and 50%, with a mean value of 36%. The greater deviation in predicting rotation may be due
to the fact that the amount of rotation is very small; therefore, the percentage deviation is high,
whereas the absolute difference between the demands from the experiments and the simulation is
not as significant. Figure 12(c) shows the comparison for peak settlement demand. It is observed
that during small shaking, the magnitude of permanent settlement is predicted reasonably well.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
694 P. RAYCHOWDHURY AND T. C. HUTCHINSON

Figure 11. (a) Moment–rotation and (b) settlement–rotation for JAU01–E08; (c) moment–rotation and
(d) settlement–rotation for JAU01–F08; (e) acceleration input time history; and (f) elastic acceleration
response spectrum of the input motion with the period of the system overlaid.

However, the simulated response of the larger footing (Station F) to the more intense earthquakes
(JAU01–F06 and JAU01–F08) over-predicts the permanent settlement by about 20%.
Comparison of the experimental and simulated residual rotations and settlements indicates
similar observations as those of the shearwall-footing models, namely that residual settlements are
reasonably predicted, whereas small residual rotations are not reasonably captured by the BNWF
model. It should be noted that measured sliding for these model tests were negligible.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
PERFORMANCE EVALUATION OF A SHALLOW FOUNDATION MODEL 695

Table V. Maximum absolute demands from experiment and simulation for bridge-footing tests.
Absolute percent
Experimental values BNWF simulation deviation
Footing |Mmax | |max | |smax | |Mmax | |max | |smax | | M | | | |s |
Event width (m) (kN m) (rad) (mm) (kN m) (rad) (mm) (%) (%) (%)
JAU01-E05 5.4 14.8 0.0030 4.2 11.1 0.0022 4.0 25 27 5
JAU01-E06 5.4 22.1 0.0093 11.0 16.6 0.0062 8.6 25 33 22
JAU01-E08 5.4 28.7 0.0220 27.0 24.7 0.0174 23.6 14 21 13
JAU01-F05 7.1 19.0 0.0010 3.0 14.1 0.0005 3.4 26 50 13
JAU01-F06 7.1 31.6 0.0040 7.0 23.7 0.0020 11.0 25 50 57
JAU01-F08 7.1 37.4 0.0088 19.0 32.5 0.0057 17.5 13 35 8
Mean 21 36 20

Figure 12. Summary of performance of the proposed BNWF model in


predicting bridge-footing test results.

4. CONCLUSIONS

Numerous experiments have provided supporting evidence that sliding, settling and rocking shallow
foundations have the potential to provide beneficial energy dissipative and period-shifting capa-
bilities to structural systems, resulting in reduction in structural force demand. However, these
beneficial attributes come at the potential expense of large transient and permanent deformations.
To this end, reliable, yet simplistic modeling tools are needed for practicing engineers to allow
them to benefit from the positive attributes of nonlinear soil–structure interaction when undertaking
a performance-based seismic design.
In this paper, a Winkler-based modeling framework to address this issue is proposed. The model
includes a distributed array of mechanistic nonlinear inelastic springs, dashpots, and gap elements,
with backbone curves of the nonlinear springs calibrated against shallow foundation experiments.
Model performance is evaluated by simulating the response of a number of centrifuge experiments.
Experiments considered include square and strip footings, bridge and building models, static and
dynamic loading, footings on sand and clay, a range of vertical factors of safety and a range of
aspect ratios. It is observed that the model can reasonably predict measured footing response in

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
696 P. RAYCHOWDHURY AND T. C. HUTCHINSON

terms of moment, shear, settlement and rotational demands. In addition, the general hysteresis
shape of the moment–rotation, settlement–rotation and shear–sliding curves is reasonably captured.
In some cases, particularly for the square footings, the moment is slightly under-predicted, which
may be due to ignoring the increase in soil capacity due to previous loading cycles and ignoring
the side friction at the front of the footing. This under-prediction is smaller when considering strip
footings because many of the footings were surface-mounted; therefore, side friction provided little
contribution to the moment capacity. The model consistently under-estimates the sliding demand.
This is perhaps due to the lack of coupling between the vertical and lateral modes of response
inherent in the model.

ACKNOWLEDGEMENTS
Support for this study was provided by the EERC Program of NSF, under Award Number EEC-9701568,
through the Pacific Earthquake Engineering Research Center (PEER). The above support is greatly
appreciated. Any opinions, findings and conclusions expressed in this paper are those of the authors,
and do not necessarily reflect those of the sponsoring organization. Suggestions from Professors Gajan,
Kutter, Stewart, Martin and Boulanger are greatly appreciated. The authors are grateful to researchers at
the University of California, Davis NEES geotechnical centrifuge; namely, Ms Key Rosebrook, Mr Jose
Ugalde, Mr Justin Phalen, and Professors Gajan and Kutter, for making available the well documented
soil-footing experimental data for use in this study.

REFERENCES
1. Housner GW. The behavior of inverted pendulum structures during earthquakes. Bulletin of the Seismic Society
of America 1963; 53(2):403–417.
2. Gadre A, Dobry R. Lateral cyclic loading centrifuge tests on square embedded footing. Journal of Geotechnical
and Geoenvironmental Engineering (ASCE) 1998; 124(11):1128–1138.
3. Maugeri M, Musumeci G, Novita D, Taylor CA. Shaking table test of failure of a shallow foundation subjected
to an eccentric load. Soil Dynamics and Earthquake Engineering 2000; 20:435–444.
4. Rosebrook KR, Kutter BL. Soil–foundation structure interaction: shallow foundations. Centrifuge Data Report
for the KRR01 Test Series: Data Report UCD/CGMDR-01/09, Center for Geotechnical Modeling, University of
California, Davis, CA, 2001.
5. Rosebrook KR, Kutter BL. Soil–foundation structure interaction: shallow foundations. Centrifuge Data Report
for the KRR02 Test Series: Data Report UCD/CGMDR-01/10, Center for Geotechnical Modeling, University of
California, Davis, CA, 2001.
6. Rosebrook KR, Kutter BL. Soil–foundation structure interaction: shallow foundations. Centrifuge Data Report
for the KRR03 Test Series: Data Report UCD/CGMDR-01/11, Center for Geotechnical Modeling, University of
California, Davis, CA, 2001.
7. Phalen J. Physical modeling of the soil–foundation interaction of spread footings subjected to lateral cyclic
loading. Master’s Thesis, University of California, Davis, CA, 2003.
8. Gajan S, Phalen J, Kutter B. Soil–foundation structure interaction: shallow foundations. Centrifuge Data Report
for SSG02: Report Number: UCD/CGMDR-03/01, Center for Geotechnical Modeling, University of California,
Davis, CA, 2003.
9. Gajan S, Phalen J, Kutter B. Soil–foundation–structure interaction: shallow foundations. Centrifuge Data Report
for SSG03: Report Number UCD/CGMDR-03/02, Center for Geotechnical Modeling, University of California,
Davis, CA, 2003.
10. Gajan S, Thomas JM, Kutter BL. Soil–foundation–structure interaction: shallow foundations. Center for
Geotechnical Modeling, University of California, Davis, CA, 2006.
11. Knappett JA, Haigh SK, Madabhushi SPG. Mechanisms of failure for shallow foundations under earthquake
loading. Proceedings of the 11th International Conference on Soil Dynamics and Earthquake Engineering, vol. 2,
University of California, Berkeley, 2004; 713–725.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
PERFORMANCE EVALUATION OF A SHALLOW FOUNDATION MODEL 697

12. Chang B, Thomas JM, Raychowdhury P, Gajan S, Kutter BL, Hutchinson TC. Soil–foundation–structure
interaction: shallow foundations. Centrifuge Data Report for the JMT02 Test Series: SSRP 07/24, University of
California, Davis, CA, 2007.
13. Ugalde JA, Kutter BL, Jeremic B, Gajan S. Centrifuge modeling of rocking behavior of structures on shallow
foundations. Proceedings of the Fourth International Conference on Earthquake Geotechnical Engineering,
Thessaloniki, Greece, 2007.
14. ATC-40. Seismic Evaluation and Retrofit of Concrete Buildings. Applied Technolgy Council (ATC), Redwood
City, CA, 1996.
15. NEHRP. Recommended Provisions for Seismic Regulations for New Buildings. Building Seismic Safety Council,
Washington, DC, 2000.
16. FEMA 356. Prestandard and Commentary for the Seismic Rehabilitation of Buildings. American Society of
Engineers, VA, 2000.
17. FEMA 440. Improvement of Nonlinear Static Seismic Analysis Procedure. Federal Emergency Management
Agency, 2005.
18. ASCE–7 05. Minimum Design Loads for Buildings and Other Structures. Structural Engineering Institute (SEI)
and American Society of Civil Engineers (ASCE), Reston, VA, 2005.
19. Gazetas G. In Foundation Engineering Handbook, Fang HY (ed.). Van Nostrand Rienhold: New York, 1991.
20. Gazetas G. The displacement and soil–structure interaction under dynamic and cyclic loading. Proceedings of
the 10th European Conference on Soil Mechanics and Foundation Engineering, Florence, 1991.
21. Taylor PW, Bartlett PE, Weissing PR. Foundation rocking under earthquake loading. Proceedings of the 10th
International Conference on Soil Mechanics and Foundation Engineering, Stockholm, Sweden, vol. 3, 1981;
313–322.
22. Chopra A, Yim SC. Simplified earthquake analysis of structures with foundation uplift. Journal of Structural
Engineering (ASCE) 1985; 111(4):906–930.
23. Yim SC, Chopra A. Simplified earthquake analysis of multistory structures with foundation uplift. Journal of
Structural Engineering (ASCE) 1985; 111(12):2708–2731.
24. Nakaki DK, Hart GC. Uplifiting response of structures subjected to earthquake motions. Report No. 2.1-3,
U.S.–Japan Coordinated Program for Masonry Building Research, Ewing, Kariotis, Englekirk and Hart, 1987.
25. Paolucci R, Pecker A. Seismic bearing capacity of shallow strip foundations on dry soils. Soils and Foundations,
Japanese Geotechnical Society 1997; 37(3):95–105.
26. Fenves GL. Effects of footing rotation on earthquake behavior of pile supported bridge piers. Technical Report
for Earth Mechanics, Inc., 1998.
27. Cremer C, Pecker A, Davenne L. Cyclic macro-element for soil–structure interaction: material and geometrical
nonlinearities. International Journal for Numerical and Analytical methods in Geomechanics 2001; 25(12):
1257–1284.
28. Allotey N, Naggar MHE. Analytical moment–rotation curves for rigid foundations based on a winkler model.
Soil Dynamics and Earthquake Engineering 2003; 23:367–381.
29. Harden CW, Hutchinson T, Martin GR, Kutter BL. Numerical modeling of the nonlinear cyclic response of
shallow foundations. Technical Report 2005/04, Pacific Earthquake Engineering Research Center (PEER), 2005.
30. Gajan S. Physical and numerical modeling of nonlinear cyclic load-deformation behavior of shallow foundations
supporting rocking shear walls. Ph.D. Thesis, University of California, Davis, CA, 2006.
31. Mylonakis G, Nikolaou S, Gazetas G. Footings under seismic loading: analysis and design issues with emphasis
on bridge foundations. Soil Dynamics and Earthquake Engineering 2006; 26:824–853.
32. Allotey N, Naggar MHE. An investigation into the winkler modeling of the cyclic response of rigid footings.
Soil Dynamics and Earthquake Engineering 2007; 28:44–57.
33. Harden CW, Hutchinson T. Beam-on-nonlinear-winkler foundation modeling of shallow, rocking-dominated
footings. Earthquake Spectra, Journal of the Earthquake Engineering Research Institute (EERI) 2009; 25(2).
34. OpenSees. Open System for Earthquake Engineering Simulation: OpenSees. Pacific Earthquake Engineering
Research Center (PEER), University of California, Berkeley, 2008. URL: http://opensees.berkeley.edu.
35. Raychowdhury P. Nonlinear winkler-based shallow foundation model for permormance assessment of seismically
loaded structures. Ph.D. Thesis, University of California, San Diego, 2008.
36. Boulanger RW. The PySimple1, TzSimple1, and QzSimple1 material models, documentation for the OpenSees
platform, 2000. URL: http://opensees.berkeley.edu.
37. Boulanger RW, Curras CJ, Kutter BL, Wilson DW, Abghari A. Seismic soil–pile–structure interaction experiments
and analyses. Journal of Geotechnical and Geoenvironmental Engineering 1999; 125(9):750–759.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe
698 P. RAYCHOWDHURY AND T. C. HUTCHINSON

38. Briaud JL, Gibbens RM. Predicted and Measured Behavior of Five Spread Footings on Sand, vol. 41. Geotechnical
Special Publication: New York, 1994.
39. Rollins K, Cole R. Cyclic lateral load behavior of a pile cap and backfill. Journal of Geotechnical and
Geoenvironmental Engineering (ASCE) 2006; 132(9):1143–1153.
40. Raychowdhury P, Hutchinson T. Nonlinear material models for winkler-based shallow foundation response
evaluation. GeoCongress 2008, Sacramento, CA, 2008.
41. Terzaghi K. Theoretical Soil Mechanics. Wiley: New York, 1943.
42. Meyerhof GG. Some recent research on the bearing capacity of foundations. Canadian Geotechnical Journal
1963; 1(1):16–26.
43. Coulomb CA. Essai sur une application des régles des maximis et minimis á quelques problémes de statique
relatifs á l’architecutre. Memoires de Mathematique et de Physique, Présentés, à l’Academie Royale des Sciences,
Paris 1776; 3:38.

Copyright q 2009 John Wiley & Sons, Ltd. Earthquake Engng Struct. Dyn. 2009; 38:679–698
DOI: 10.1002/eqe

You might also like