Anda di halaman 1dari 17

INTERFACE MODEL FOR DYNAMIC

SOIL-STRUCTURE INTERACTION
By M d . Musharraf-uz Z a m a n , 1 A. M . ASCE,
C h a n d r a k a n t S. Desai, 2 M . ASCE, a n d
Eric C. D r u m m , 3 A. M . ASCE
Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

ABSTRACT: A simple thin-layer element is developed and u s e ' in a finite ele-


ment procedure for simulation of various modes of deformanon in dynamic
soil-structure interaction. The constitutive behavior of the interface is defined
by decomposing it into its normal and shear components. The soil is modeled
as an elastic-plastic hardening material. The numerical procedure is used to
predict behavior of a model structure tested in the field, and the influence of
interface behavior on displacements, velocities, and accelerations is delineated.

INTRODUCTION

The response of a structure-foundation system subjected to dynamic


loadings such as those resulting from earthquakes and blasts can be in-
fluenced significantly by the characteristics of interfaces or junctions be-
tween the structure and foundation (Fig. 1). Often, a design analysis of
structures and foundations subjected to dynamic loads is performed by
assuming complete bonding at the interfaces at all stages of loading.
Although that assumption usually simplifies an analysis procedure sig-
nificantly, it can account for soil-structure interaction effects only to a
limited extent, because the relative motions are not included in the anal-
ysis. Under dynamic loadings, relative movements such as sliding, sep-
aration or debonding can occur at interfaces at different stages of load-
ing. Thus, for a realistic analysis of these problems, it may be appropriate
to incorporate such motions at the interfaces.
Incorporation of deformation behavior of interfaces in a dynamic soil-
structure interaction problem can introduce strong nonlinearity in the
system. Analytical solution of those problems can be conceptually and
mathematically difficult. As a result, numerical methods, such as the
finite element method, have often been used.
Review.—Although the importance of interface behavior and the re-
sulting soil-structure interaction effects have been recognized for a long
time, development of appropriate models to simulate them has been un-
dertaken only recently. In the recent past, a number of models have
been reported in the literature for simulating interface (and joint) be-
havior under dynamic loadings. Although the writers do not present a
detailed review on existing models, comprehensive reviews on the sub-
ject can be found in various publications (2,3,10,21). Only a few recent
lAsst. Prof., Dept. of Civ. Engrg., Univ. of Oklahoma, Norman, Okla.
2
Prof., Dept. of Civ. Engrg. and Engrg. Mechanics, Univ. of Arizona, Tucson,
Ariz.
3
Asst. Prof., Dept. of Civ. Engrg., Univ. of Tennessee, Knoxville, Tenn.
Note.—Discussion open until February 1, 1985. To extend the closing date one
month, a written request must be filed with the ASCE Manager of Technical and
Professional Publications. The manuscript for this paper was submitted for re-
view and possible publication on October 7, 1983. This paper is part of the Jour-
nal of Geotechnical Engineering, Vol. 110, No. 9, September, 1984. ©ASCE, ISSN
0733-9410/84/0009-1257/$01.00. Paper No. 19133.
1257

J. Geotech. Engrg., 1984, 110(9): 1257-1273


Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

FIG. 1.-—Soil-Structure Interaction Problems: (a) Structure-Foundation System; (b)


Retaining Wall

works related directly to this study are reviewed here.


Toki, Sato, and Miura (17) used the interface element proposed by
Goodman et al. (9) to study the behavior of a structure-foundation sys-
tem under dynamic loads. The separation and sliding aspects of inter-
face behavior were included in the formulation, and the method was
applied to: (1) A model nuclear reactor building resting on the surface
of a layered soil medium; and (2) a partially embedded pier foundation
structure subjected to dynamic excitations. It was observed that, assum-
ing perfect bond at the contact surface between soil and structure, the
structure's motion was restricted by the surrounding subsoil, thus
underestimating the actual response of the structure. Since no trans-
mitting boundaries were used and a high value of damping was adopted,
these results may not represent practical situations.
Isenberg and Vaughan (12) and Vaughan and Isenberg (18,19) gave a
detailed consideration to interface behavior under dynamic loading. A
continuum (soil) element was used as an interface element by assigning
special properties to it. Sliding, separation or debonding and rebonding
aspects were simulated by imposing constraints upon their allowable stress
states in addition to those imposed by the constitutive relations for soil.
These additional constraints restrict the normal and shear forces trans-
ferred between soil and structure. The model was used to obtain solu-
tions for a number of interaction problems including nuclear contain-
ment structures and buried cylinders subjected to shock loading from a
traveling air blast. A number of other investigators (8,11,13,14,20-22)
have also considered interface behavior in (dynamic) soil-structure
interactions.
Objectives.—The objectives of this paper are: (1) To present a sim-
plified model, called thin-layer element, for simulation of interface be-
havior under dynamic loading; and (2) to verify the proposed model by
comparing predictions with field observations of the response of a model
nuclear containment structure due to SIMQUAKE excitation (12,18,19).

MODES OF DEFORMATION AT INTERFACE

As mentioned previously, interfaces or junctions between structures


and foundations can experience relative motions under dynamic load-
ings. These relative motions include translational, rotational, and rock-
1258

J. Geotech. Engrg., 1984, 110(9): 1257-1273


Body 1 ! Body 1
! A'

8ody 2 Body 2
Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

77?—w—^—sr" ^e ^—3e ?3*r

C:
/ Body 1
M" r
A A
" c 4> A

Body 1
c -I

A = Total Area

Body 2 A, = Contact Area Body 2

•vfe—TJP—^—?m 7W 1
=75 >SS *W

FIG. 2.—Modes of Deformation at Interface: (a) Stick or No-Slip; (b) Slip; (c) De-
bonding; (d) Rebonding

ing motions. There are 4 basic modes of deformation that an interface


element can undergo: (1) Stick or no-slip; (2) slip or sliding; (3) sepa-
ration or debonding; and (4) rebonding. Fig. 2 shows various modes of
deformation for 2-dimensional idealization. An interface element is said
to be in stick or no-slip mode when there is no relative movement be-
tween the adjoining bodies. Slip or sliding occurs when relative move-
ments take place in such a manner that the contact between the mating
bodies is maintained. Separation or debonding mode occurs when gaps
open up between two bodies that were in contact previously. An inter-
face element in separation mode can return to stick mode in subsequent
loading, which is referred to as rebonding. Details of simulation of these
modes are given here.

THIN-LAYER ELEMENT

Fig. 3 schematically shows the proposed thin-layer element for sim-


ulation of interface and joint behavior (2,7). This element has been used
successfully for solution of a number of static 2- and 3-dimensional soil-
structure interaction problems, such as rail-track support structures, buried
pipes, retaining walls, and tunnels (4,5,15,16). In this study, the thin-
layer element is modified and implemented for dynamic soil-structure
interaction problems. This procedure involves use of the displacement
finite element procedure, in which equations of the thin-layer element
are derived separately and then added to those for the soil and structural
elements.
The underlying idea of the thin-layer element is based on the as-
sumption that the behavior near the interface (Fig. 1) involves a finite
thin zone (Fig. 3), rather than a zero thickness zone as assumed in sev-
eral previous investigations (9,10,17). The behavior of this thin zone or
layer can be significantly different from the behavior of the surrounding
structural and geological materials. This behavior is achieved by adopt-
1259

J. Geotech. Engrg., 1984, 110(9): 1257-1273


Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

B = average width of
side of element

FIG. 3.—Thin-Layer (Interface) Element

ing appropriate constitutive law for the element. In other words, the
thin-layer element is treated essentially like any other solid (soil, rock
or structural) element, but its constitutive relations are denned differently.
Constitutive Relation for Thin-Layer Element.—The constitutive re-
lation matrix, [C,], of the thin-layer element is expressed as
L*" nn J i L*-* ns J i
[C,] = (i)
-\S-$n\i V-ss\i.
in which, [C„„] = normal component; [Css] = shear component; and [Cm],
[Csn] represent coupling effects between normal and shear behaviors. At
this time, the coupling components are not considered. The subscript,
i, denotes that the quantities are related to interface.
Normal Behavior.—In most of the previous models related to the be-
havior of interface, a high value was assigned for the normal sitffness
to avoid interpenetration of neighboring solid elements. Selection of this
high value has often been arbitrary. Since the interface is surrounded
by the structural and geological materials, its normal properties during
the deformation process should be dependent upon the characteristics
of the thin interface zone, as well as the state of stress and properties
of the surrounding elements. Based on these considerations, it was pro-
posed to express the normal stiffness as (2)
[C„„],- = [C„„(aL,Pxm,7m)] (2)
in which a'm/ (3f,, y„ (m = 1, 2, ...) = the properties of the interface,
geological (soil or rock), and structural elements, respectively. For nu-
merical application, Eq. 2 needs to be written explicitly in a suitable
functional form as
[C„„]i = Xi[C„„], + A.2[C„„]g + X.3[C nn\st • • (3)
in which [C„„], = normal behavior of thin layer; subscripts g and sf =
geological and structural materials, respectively, and Xx, X.2 / a n d X3 are
the participation factors, which may vary from 0 to 1. One of the sim-
plifications of Eq. 3 would be to assume X2 = ^ 3 = 0 and \ 2 = 1, implying
that the normal component is based on the normal behavior of the thin-
layer element alone, evaluated just as the adjacent soil element. It can
be possible to arrive at appropriate values for \ , based on a trial and
1260

J. Geotech. Engrg., 1984, 110(9): 1257-1273


error procedure in which numerical solutions are compared with labo-
ratory or field observations. It was found that often satisfactory results
can be obtained by assigning the same properties as the geologic ma-
terial for the interface normal behavior; the results presented herein are
based on this assumption.
Shear Behavior.—Shear part, [C ss ],, of the interface constitutive re-
lation matrix is obtained using the results of direct shear tests (static or
Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

cyclic) of interfaces. The procedure for evaluation of [C ss ], is shown in


Fig. 4. The relative (small) shear displacement, u°, is related to the shear
strain, 0 by

9 3
> <4>
in which t' = thickness of the interface element; and the superscript, 0,
= total quantity. Relating 6 to the tangent shear modulus, G,, of the
interface yields

G, = (5)

in which T° = the total shear stress at the interface. For 2-dimensional


(plane strain) problems, the size of [Css], is 1 x 1, which is the shear
modulus, G,, of the interface. In a soil-structure interaction problem,
the value of G, may depend on several factors, such as amplitude of
relative displacement, u°; interface thickness, t'; shear stress, T°; normal
stress, CT°„; and interface roughness, \L. In the case of cyclic loading, G,
is also dependent on number of cycles, N. Thus, G, is expressed as
G, = F(<T0nn,r0,u°r,N,t') (6)
Implementation of this model in a finite element procedure will require

Body 1

Body 2
Interface

(a) Schematic of Direct Shear Test

H
(b) Deformations at the Interface

FIG. 4.—Shear Behavior of Interface: (a) Direct Shear Test; (b) Deformations at
Interface
1261

J. Geotech. Engrg., 1984, 110(9): 1257-1273


an explicit expression for Eq. 6. Any logical expression can be used for
this purpose as long as the related coefficients of the function can be
evaluated from laboratory tests. Here, a polynomial form is used:
0\2
T° = a! + a2u° + a3(u°j- (7)
where, a, = /,(aS„,N); (z = 1, 2, 3). (8)
Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

in which a x , a2, and a 3 = coefficients or interface parameters, a, (i =


1, 2, 3) can be expressed in the following form:
«, = (Pi), + (P2),cx°„ + ((J 3 ),(<T°„) 2 + (P 4 ),N + (p 5 ),N 2 (9)
Coefficients a, and pf (i = 1, 2, 3, ...) can be evaluated from appropriate
laboratory tests.
By assuming a°nn and N to be constant at a given instant of time, and
V as constant, G, can be approximated by using Eqs. 4-5 as

G,= (10)
du° o-f}„,N=constant

Eq. 10 can be used to estimate G, for elements in stick and sliding modes.
It may be noted here that several previously developed models (9,10,17)
assumed arbitrary and small values for shearing modulus, G,, for ele-
ments in sliding mode, and often, selection of this (low) value did not
have a logical basis. In the present formulation, this assumption is not
necessary. Use of appropriate values for coefficients a, and p, (Eq. 9)
automatically yields small values for G, under sliding modes.
Another important feature incorporated in the thin-layer element is
loading, unloading, and reverse loading behavior of interfaces (Fig. 5).
The loading and reverse loading portions are simulated by using Eq. 10,
whereas a constant modulus is used for the unloading behavior. It may
be also noted that when separation or debonding occurs, the shear stiff-
ness of the interface tends toward zero. This aspect is incorporated in

Unloading

U - constant

FIG. 5.—Typical Loading, Unloading, and Reloading Curves


1262

J. Geotech. Engrg., 1984, 110(9): 1257-1273


the computer code by identifying elements that are undergoing sepa-
ration and by reducing their moduli to small values, about 1/10 of the
initial value.
As can be seen from Eq. 10, the shearing modulus, G,, is dependent
on the thickness of the thin-layer element, t', which in turn can influ-
ence the quality of simulation of interface behavior. Solution of a num-
ber of static and dynamic interaction problems (7,23) has indicated that
Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

satisfactory simulation can be obtained for t'/B ratios (Fig. 3) in the range
of 0.01-0.1 Here, B is the average width of the adjoining elements.
Evaluation of a, and p;.—As mentioned previously, coefficients a, and
P; associated with the shear behavior of thin-layer element can be de-
termined from the results of appropriate laboratory tests. In this study,
a series of cyclic tests were performed with interfaces between Ottawa
sand and concrete, by using a cyclic multi-degree-of-freedom shear de-
vice (2,6). This device can be essentially viewed as a large direct shear
device in which tests can be performed under static or cyclic loading.
To perform a cyclic test, a normal load is applied first to simulate in situ
stress, and then a cyclic displacement or shear load is applied. Average
interface shear stress, T°, and relative displacement, u°, histories are
recorded at selected number of cycles, N. Fig. 5 schematically shows T°
versus u°r. The coefficients a, and p, (Eqs. 7 and 9) are determined from
such plots by using a curve fitting procedure. Typical values of these
coefficients obtained from sand-concrete interface tests are given in Ta-
ble 1. As an approximation, these values were used to analyze the model
structure described here.

SIMULATION OF DEFORMATION MODES

Various modes of deformation outlined previously are incorporated in


the thin-layer element by using appropriate stress redistribution schemes.
It is assumed that all interface elements are initially in the stick mode,
and excitations are applied incrementally with a time increment of Ai.
Based on the configuration of the system at time t, a trial response at t
+ Afis evaluated by solving the global equations of dynamic equilibrium
(Eq. 17). The following procedure is then adopted to check if correction
of the computed trial response is required:

TABLE 1.—Values of Coefficients


Values3 of Coefficients in Eq. 9

Pi <*i <x2 ot3


(1) (2) (3) (4)
p! 0.635 143.03 -287.94
P2 0.0207 4.70 -9.645
P3 -0.0005 -0.12054 0.253
P4 0.038 8.937 -18.66
P5 -0.000088 00.02235 -0.04587
"Units are in lb and in.
Note: 1 lb = 4.448 N; 1 in. = 2.54 cm.

1263

J. Geotech. Engrg., 1984, 110(9): 1257-1273


1. Previous mode = stick or no-slip
(a) Known: « ) ' , (x°)f, (u°„)f;
(b) Trial response: Aa„„, AT and &u„

in which u„ = average normal relative displacement; A = an incremental


quantity; and t = time level. All variables are referred to a local coor-
dinate system of an interface element. The element will stay in stick
Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

mode if
{(a°„)' + A<r„„} > 0 (compressive) (11)
and Fs = (sgn) {(T°)< + AT} - [c. + {(«r°,)' + Acr„„} tan 8] < 0 (12)
in which ca = activated adhesion; 8 = activated friction angle; and (sgn)
= 1 if {(T 0 )' + AT} > 0, and (sgn) = - 1 if {(T 0 )' + AT} < 0. Fs in Eq. 12
is called a slip function, and is shown in Fig. 6. The element can go into
slip mode if {(cr°„ + ACT„„} > 0 (compressive)
and Fs = (sgn) {(T 0 )' + AT} - [c. + {(*"„)' + A<r„„} tan 8] > 0 (13)
Iterations are not required if Fs = 0. Otherwise, iterations are performed
within the time step, At, until Fs ss 0 is satisfied. At each iteration, the
residual value of Fs, which has the dimension of stress, is converted
into forces and applied to the system as self-equilibrating nodal loads.
The element can go into separation or debonding mode if
{ « ) ' + Aa„„} < 0 (tensile) (14)
In the separation mode, neither normal nor shear traction can be trans-
mitted through the interface. Assume that for an element in the sepa-
ration mode, it is found that
{(<T°m)< + Acr„„} = R, (15)
0
and {(T )' + AT} = R2 (16)

Iterations are not required if R1 - i?2 = 0. Otherwise, residuals Ki and


R 2, which have the dimension of stress, are converted into self-equili-
brating nodal loads that are applied to the system during the iterative
corrections.
Rebonding is assumed to occur if

Slip Cone

FIG. 6.—Graphical Representation of Slip Function, Fs


1264

J. Geotech. Engrg., 1984, 110(9): 1257-1273


{(cr°„)' + Acr„„}>0 (17)
and Fs < 0 , (18)
As an approximation, full contact over an element is assumed during
rebonding.
With the thin-layer element, generally the adjacent nodes do not pen-
etrate each other. However, if penetration occurs, it is prevented by sep-
Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

arating a given pair of adjacent nodes by an arbitrary fraction, e, of the


thickness, t', of the element. The magnitude of the penetration is found
on the basis of the computed values of normal displacements of the nodes.
Then, the value of the normal displacement to be suppressed is found
as the difference between the penetration and the value of separation to
be established, equal to ef'. The magnitude of the suppressed displace-
ment, {A</s}, is converted into an equivalent correction load vector, {AQ},
as

{AQ} = I [C]{ [B]{Aqs}dV (19)


Jv
in which [C]{ = constitutive matrix at time, t, and iteration z; [B] = trans-
formation matrix; and V = element volume. The iterative procedure is
continued until convergence, defined by \Aq3\ < X, in which X = a small
number.
It should be mentioned that global stiffness, mass, and damping ma-
trices (Eq. 20) are formed once during each time step; only the load vec-
tor is changed during the iterations. Also, velocity and acceleration vec-
tors are not updated during iterations.
Finite Element Procedure.—The formulation of the thin-layer element
outlined in the preceding sections has been incorporated in a displace-
ment finite element procedure for solving dynamic soil-structure inter-
action problems with 2-dimensional idealization in which the structural,
soil, and interfaces are modeled using 8-noded isoparametric elements.
The equations of global dynamic equilibrium are obtained by adding
the contribution of the interface elements to those of the soil and struc-
tural elements. The final equations are expressed as
[M]{q} + [C]{q} + [K]{q} = {Q«>} (20)
in which [M], [C], and [K] = global mass, damping, and stiffness matri-
ces, respectively; {Q(t)} and {q} = global load and displacement vectors,
respectively; and the overdot denotes derivative with respect to time.
The matrix equations in Eq. 20 are solved by using Newmark's Beta
method with coefficients 7 = 1/2 and (3 = 1/4.
Nonlinear Behavior.—The structural elements are assumed to be lin-
early elastic, whereas the soil can be assumed linearly elastic or elastic-
plastic. For the latter, a hardening cap model is used (12,18) (see Fig.
7). The equations for the failure (fixed) surface, cap or yield surface, and
hardening are given as
F,(h,y/hD) = Vhv - [« + 6/t - 7 e x p ( - B / 0 ] (21a)
Fc = R2J2D ~ ( X - Lf + (/x - L)2 (21b)
1265

J. Geotech. Engrg., 1984, 110(9): 1257-1273


tf t
Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

, Z _ ^ L \.— Rb -|x

FIG. 7.—Cap Model

(eg) = W[l - exp (-DX)] (21c)


=
in which ]\ = first invariant of stress tensors; }2D second invariant of
the deviatoric stress tensor; and a, p, y, 0, D, and W = material param-
eters shown in Fig. 7.

VERIFICATION

The accuracy and performance of the proposed thin-layer element have


been verified with respect to a number of static and dynamic soil-struc-
ture interaction problems in which predictions have been compared with
observations, previous numerical procedures, and closed-form solu-
tions, wherever available. These static problems include: (1) Behavior of
a buried pipe and tunnels (4,7,13,23); (2) behavior of a strutted retaining
wall at various stages of construction (5,15,16); (3) simulation of static
direct shear tests with interfaces between concrete and sand (7); and (4)
behavior of uniformly loaded circular rafts resting in frictional contact
with elastic half space (23). Some of the dynamic soil-structure interac-
tion problems that were investigated to verify the proposed model in-
clude: (1) Simulation of cyclic shear tests with interfaces between Ottawa
sand and concrete (23); (2) behavior of a model nuclear reactor building
and a pier foundation structure subjected to cyclic excitation (17,23); and
(3) simulation of behavior of a model nuclear containment structure sub-
jected to SIMQUAKE loading (12,18,23). Satisfactory and consistent re-
sults were obtained in all these cases.
In addition to the verification of the proposed thin-layer element for
the foregoing static and dynamic problems, it was verified with respect
to a general interface element (23,24). Here, in contrast to the proposed
approach of adding interface equations to those for soils and structures
derived from the displacement procedure, a mixed procedure was used.
It involved expressing the variational function as composed of the po-
tential and kinetic energy terms, as in the displacement method, and
energy terms due to constraint conditions of the interface. This proce-
dure resulted in a coupled set of equations for the structure, soil, and
1266

J. Geotech. Engrg., 1984, 110(9): 1257-1273


interfaces. Predictions from both approaches were compared with each
other for the problem of an underground pipe subjected to overburden
pressure. Both approaches provided very good comparisons with closed-
form and other finite element solutions (13). However, the thin-layer
element proposed herein was found to be simpler and more economical
compared to the general mixed element. Also, it was found to be as
good as or better than the conventional zero-thickness element (9), par-
Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

ticularly in accounting for debonding and rebonding modes and in sim-


ulating interfaces in flexible structures and soils (5,7,15). Thus, the thin-
layer element is used in this study. Results of its implementation for a
typical example involving dynamic analysis are described here.

EXAMPLE: MODEL NUCLEAR CONTAINMENT STRUCTURE

The field test program, known as SIMQUAKE, was conducted by the


University of New Mexico under a research project sponsored by the
Electric Power Research Institute (EPRI), in which response of model
nuclear containment structures due to strong ground motion (generated
by blasts) was recorded. Details of the tests and measurements are given
by Vaughan and Isenberg (18,19). The response of the 1/8 scale model
nuclear containment structure is analyzed here using a plane strain finite
element idealization. The finite element mesh used in the simulation is
shown in Fig. 8. The material properties used in the analyses are given
in Tables 1-2. Interface elements are assigned the same material prop-
erties for normal behavior as the properties in normal behavior for the
adjoining soil elements.
Excitation.—During the SIMQUAKE test, ground response was re-
corded at selected points along the boundaries of the soil island (18,19).

TABLE 2.—Cap Model Parameters for Materials for Soils and Elastic Parameters
for Concrete (18,19)
Parameters Soil 1 Soil 2 Backfill Concrete
(1) (2) (3) (4) (5)
E psi 46,435 22,113 13,100 4 x 106
(kPa) (3.2 x 105) (1.5 X 105) (9.0 x 104) (2.7 X 107)
V 0.3 0.3 0.26 0.2
a psi 470.0 470.0 470.0
(kPa) (3,240.2) (3,240.2) (3,240.2)
e 0.0 0.0 0.0
P Vpsi 0.165 0.165 0.165
(1/kPa) (0.024) (0.024) (0.024)
7 psi 390.0 390.0 390.0
(kPa) (2,688.7) (2,688.7) (2,688.7)
D 1/psi 0.0007 0.0007 0.00065
(1/kPa) (0.000102) (0.000102) (0.000094)
W 0.06 0.06 0.06
z 0.0 0.0 0.0
R 2.5 2.5 2.5
-yw lb/in. 3 0.0637 0.0637 0.0637 0.058
(kg/cm3) (0.00176) (0.00176) (0.00176) (0.0016)

1267

J. Geotech. Engrg., 1984, 110(9): 1257-1273


••* V i s & l Backfill E D Soil 1

:
;;V.\;>v,: *7 :'•''-'$& V'-i v; : '?.*:U-
U •ilfi KKVK A
Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

• ; • ! •

^ i

; - , ••',

• ; : . • • . ' '
w
FIG. 8.—Mesh Used in Simulation of Soil-Structure Interaction Due to SIMQUAKE
II

I 5-0

FIGS. 9.—Selected Horizontal Displacement-Time History at Two Points Used as


Input for Analyses of SIMQUAKE II

The velocity-time histories were provided by the Weidlinger Associates,


Menlo Park, Calif. They were integrated numerically with At - 0.001 to
obtain displacement-time histories at various points of measurements on
the boundaries of the soil island (18). The displacement-time histories
(Fig. 9) at nodal points on the boundaries of the finite element grid (Fig.
8) were obtained by linear interpolation based on the foregoing displace-
ment histories.
Prediction of Structural Response.—Fig. 10 shows the computed hor-
izontal velocity-time histories at the top of structure for both the fric-
tional and bonded interface conditions. The measured velocity history
shown in this figure is reproduced from Ref. 18. The frictional interface
case allows for the relative motions discussed earlier, whereas the bonded
case does not permit interface motions. The correlation between com-
1268

J. Geotech. Engrg., 1984, 110(9): 1257-1273


Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

-100.0 ' ' 1 '


0.0 1.0 2.0 3.0
Time (sec)

FIG. 10.—Comparison of Computed and Measured Horizontal Velocity-Time His-


tory at Top of Structure (Point P)

puted and measured response is found to be very good for the first cycle
of strong shaking, approximately from t = 0.0-0.8 s. During this cycle,
the modeling of interface behavior as frictional or bonded does not sig-
nificantly influence the response. For the second phase of strong mo-
tion, the computed values underestimate the measured response. This
difference is particularly noticeable after t = 1.35 s. However, the overall
correlation given by the frictional interface condition is considered to be
better. After t = 1.8 s, the peak response from the bonded case is lower
in magnitude than that from the frictional case. However, after t =
2.2 s, the bonded case yields higher magnitude of the response.
Fig. 11 shows the computed horizontal displacement history at the top
of structure for both bonded and frictional interface conditions. Since
measurements were not available, only computed values are compared.
It can be noticed that for the horizontal displacement, the bonded con-
dition predicts the lower response. Particularly for t > 1 s, the horizontal
response is underestimated by the bonded interface assumption.
Distribution of Contact Stress.—In the SIMQUAKE test, normal con-
tact stress histories at selected points on the interface region were re-
corded (18). A typical comparison between the calculated and measured
contact stress histories is presented in the following description.
Fig. 12 compares measured and calculated normal contact stresses be-
neath the bottom left corner of the structure for frictional and bonded
interface conditions. In the finite element simulation with the frictional
interface (dynamic), normal tensile stress is not allowed to exceed 10.5
psi (72.4 kPa). This value corresponds to initial static stresses induced
by the self-weight of the structure. It is observed that computed and
measured stresses for the frictional case correlate well up to t = 1.0 s.
After 1.0 s, the predicted values differ significantly from the measured
values in terms of magnitude as well as phase. Peak (compressive) com-
puted responses are generally higher than the measured response. A

1269

J. Geotech. Engrg., 1984, 110(9): 1257-1273


Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

- Fractional Interface
-Bonded Interface
1 inch = 2.54 cm

Time (sec)

FIG. 11.—Influence of Interface Condition On Horizontal Displacement-Time His-


tory at Top of Structure (Point P)

FIG. 12.—Comparison of Computed and Measured Normal Contract Stress-Time


Histories at Bottom of Structure (Point Q)

reverse trend is observed for tensile contact stress. The computed in-
terface stresses represent average stresses at integration points of an in-
terface element. Thus, the computed and measured stresses do not ex-
actly correspond to same location. The difference between the computed
and predicted responses can be partly attributed to this reason. The pre-
dictions from the bonded case are quite erratic and show significant dif-
ferences in phase. In general, predictions by frictional interface condi-
1270

J. Geotech. Engrg., 1984, 110(9): 1257-1273


tion appear to be better than those from the bonded case.
Comments.—The provision of interface motions can increase or de-
crease the response (displacements, velocities, accelerations) depending
upon the location, time period, frequency, and other factors related to
loading and physical properties. This response is indicated in the pre-
ceding field problem and in a number of finite element predictions (23);
similar mixed trends have also been reported previously by Wolf (22).
Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

Based on the comparisons for the field problem and the verification
the writers believe that the proposed interface model can provide a sat-
isfactory and consistent representation of the interface behavior.

CONCLUSIONS

A simple thin-layer element that allows for deformation modes, such


as no slip, slip, debonding and rebonding, is used for simulating inter-
face behavior between structural and geologic media subjected to dy-
namic loading. Predictions from the finite element procedure that in-
corporates the thin-layer element compared well with available solutions
and observations of the response of a model nuclear power plant struc-
ture. Based on this study, it is believed that the proposed element can
provide satisfactory and consistent formulation of interface behavior un-
der dynamic loading. In view of the complexity of the problem and in-
fluence of a number of factors, such as geometry (including reflections
and refraction at discretized boundaries), type of loading, material prop-
erties, time integration, and mesh layouts, further investigations will be
needed in order to delineate their effects on the dynamic response.

ACKNOWLEDGMENTS

The results presented herein were obtained from research conducted


under Grant No. 81-CE-94 from Earthquake Hazards Mitigation Pro-
gram, National Science Foundation, Washington, D.C. Sincere appre-
ciation is due to J. Isenberg and D. K. Vaughan, Weidlinger Associates,
Menlo Park, Calif., for providing the velocity-time histories for the
SIMQUAKE test and for useful discussions. Permission was granted by
Electric Power Research Institute, Menlo Park, Calif., to release the ve-
locity-time histories. Useful discussions with J. T. Christian, Stone &
Webster/Boston, Mass., are gratefully acknowledged.

APPENDIX.—REFERENCES

1. Bathe, K. J., and Wilson, E. L., Numerical Methods in Finite Element Analysis,
Prentice-Hall, Inc., Englewood Cliffs, N.J., 1976.
2. Desai, C. S., "Behavior of Interfaces Between Structural and Geologic Me-
dia," International Conference on Recent Advances in Geotechnical Earthquake En-
gineering and Soil Dynamics, St. Louis, Mo., 1981.
3. Desai, C. S., and Christian, J. T. (eds.), Numerical Methods in Geotechnical En-
gineering, McGraw-Hill Book Co., Inc., New York, N.Y., 1977.
4. Desai, C. S., Eitani, I. M., and Haycocks, C , "An Application of Finite Ele-
ment Procedure for Underground Structures with Nonlinear Materials and

1271

J. Geotech. Engrg., 1984, 110(9): 1257-1273


Joints," Proceedings of the Fifth International Congress of the Society of Rock Me-
chanics, Melbourne, Australia, Apr., 1983.
5. Desai, C. S., Lightner, J. G., and Sargand, S. M., "Mixed and Hybrid Pro-
cedures for Nonlinear Problems in Geomechanics," Proceedings of the Fourth
International Conference on Numerical Methods in Geomechanics, University of Al-
berta, Edmonton, Canada, 1982.
6. Desai, C. S., "A Dynamic Multi-Degree-of-Freedom Shear Device," Report
No. 80-36, Department of Civil Engineering, Virginia Technical College,
Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

Blacksburg, Va., 1981.


7. Desai, C. S., Zaman, M. M., Lightner, J. G., and Siriwardane, H. J., "Thin-
Layer Elements for Interfaces and Joints," International Journal for Numerical
and Analytic Methods in Geomechanics, Vol. 8, No. 1, 1984, pp. 19-43.
8. Ghaboussi, J., Wilson, E. L., and Isenberg, J., "Finite Element for Rock Joint
and Interfaces," Journal of the Soil Mechanics and Foundations Division, ASCE,
Vol. 99, No. SM10, pp. 833-848.
9. Goodman, R. E., Taylor, R. L., and Brekke, T. L., "A Model for the Me-
chanics of Jointed Rock," Journal of the Soil Mechanics and Foundation Division,
ASCE, Vol. 94, No. SM3, 1968, pp. 637-659.
10. Heuze, F. E., and Barbour, T. G., "New Models for Rock Joints and Inter-
faces," Journal of the Geotechnical Engineering Division, ASCE, Vol. 99, No. TE4,
1973, pp. 887-908.
11. Isenberg, J., Lee, L., and Agbabian, M. S., "Response of Structures to Com-
bined Blast Effects," Journal of the Transportation Engineering Division, ASCE,
Vol. 99, No. TE4, 1973, pp. 887-908.
12. Isenberg, J., and Vaughan, D. K., "Nonlinear Effects in Soil-Structure Inter-
action," Proceedings of the Symposium on Implementation of Computer Procedures
and Stress-Strain Laws in Geotechnical Engineering, held at Chicago, 111., 1981,
pp. 29-44.
13. Katona, M. G., "A Simple Contact-Friction Interface Element with Applica-
tions to Buried Culvert," Proceedings of the Symposium on Implementation of
Computer Procedures and Stress-Strain Laws in Geotechnical Engineering, Chicago,
111., Vol. 1, 1981.
14. Kausel, E., et al., "Seismically Induced Sliding of Massive Structures," Jour-
nal of the Geotechnical Engineering Division, ASCE, Vol. 15, No. GT12, 1979,
pp. 1471-1488.
15. Lightner, J. G., "A Mixed Finite Element Procedure for Soil-Structure Inter-
action Including Construction Sequences," thesis presented to the Virginia
Polytechnical Institute and State University, at Blacksburg, Va., in 1981, in
partial fulfillment of the requirements for the degree of Doctor of Philosophy.
16. Siriwardane, H. J., "Nonlinear Soil-Structure Interaction Analysis of One-,
Two- and Three-Dimensional Problems Using Finite Element Method," the-
sis presented to the Virginia Polytechnical Institute and State University, at
Blackburg, Va., in 1980, in partial fulfillment of the requirements for the de-
gree of Doctor of Philosophy.
17. Toki, T., Sato, T., and Miura, F., "Separation and Sliding Between Soil and
Structure During Strong Ground Motion," Earthquake Engineering and Struc-
tural Dynamics, Vol. 9, 1981, pp. 263-277.
18. Vaughan, D. K., and Isenberg, J., Nonlinear Soil-Structure Analysis of SIMQUAKE
II, Research Project 810-2, Final Report for Electric Power Research Institute
(EPRI), 1982.
19. Vaughan, D. K., and Isenberg, J., "Nonlinear Rocking Response of Model
Containment Structures," Earthquake Engineering & Structural Dynamics, Vol.
11, 1983, pp. 275-296.
20. Wolf, J. P., "Soil-Structure Interaction with Separation of Base Mat From Soil,"
Nuclear Engineering and Design, Vol. 38, 1976, pp. 357-384.
21. Wolf, J. P., "Seismic Response Due to Travelling Shear Wave Including Soil-
Structure Interaction With Base-Mat Uplift," Earthquake Engineering and Struc-
tural Dynamics, Vol. 5, 1977, pp. 337-363.
22. Wolf, J. P., "Response of Structures Permitting Lift-Off to Rotational Input
1272

J. Geotech. Engrg., 1984, 110(9): 1257-1273


Motion From Horizontally Propagating Waves," Proceedings, Second ASCE
Conference on Civil Engineering and Nuclear Power, Vol. VI, Knoxville, Tenn.,
Sept., 1980.
23. Zaman, M. M., "Influence of Interface Behavior in Dynamic Soil-Structure
Interaction Problems," thesis presented to the University of Arizona, at Tuc-
son, Ariz., in 1982, in partial fulfillment of the requirements for the degree
of Doctor of Philosophy.
24. Zaman, M. M., and Desai, C. S., "Models for Sliding and Separation of In-
Downloaded from ascelibrary.org by Ryerson University on 01/05/19. Copyright ASCE. For personal use only; all rights reserved.

terfaces Under Static and Cyclic Loading," Proceedings, International Conference


of Constitutive Laws for Engineering Materials, Tucson, Ariz., 1983.

1273

J. Geotech. Engrg., 1984, 110(9): 1257-1273

Anda mungkin juga menyukai