You are on page 1of 9

Nonlinear Model for Lead–Rubber Bearings Including

Axial-Load Effects
Keri L. Ryan, M.ASCE1; James M. Kelly, M.ASCE2; and Anil K. Chopra, M.ASCE3

Abstract: Existing models for isolation bearings neglect certain aspects of their response behavior. For instance, rubber bearings have
been observed to decrease in stiffness with increasing axial load, and soften in the vertical direction at large lateral deformations. The yield
Downloaded from ascelibrary.org by Iowa State University on 09/27/13. Copyright ASCE. For personal use only; all rights reserved.

strength of lead–rubber bearings has also been observed to vary with axial load, such that a lightly loaded bearing may not achieve its
theoretical strength. Models that include these axial-load effects in lead–rubber bearings are developed by extending an existing linear
two-spring model to include nonlinear behavior. The nonlinearity includes an empirical equation for the experimentally observed variation
of yield strength. For numerical implementation, the bearing forces are found by solving the nonlinear equilibrium and kinematic
equations using Newton’s method, and the instantaneous bearing stiffness matrix is formed from the differentials of these equations. The
response behavior of the models is confirmed by comparison with experimental data.
DOI: 10.1061/共ASCE兲0733-9399共2005兲131:12共1270兲
CE Database subject headings: Axial forces; Axial loads; Rubber; Lead; Base isolation; Numerical models; Nonlinear analysis;
Plasticity.

Introduction ception, some HDR bearings showed increased stiffness or hard-


ening at large shear strains that was magnified by large axial loads
Base isolation is a seismic technique that has been traditionally 共Aiken et al. 1989兲. The stiffness reduction appeared to be great-
reserved for short or squat structures. However, isolation of taller est in bearings with low shape factors.
buildings is becoming more common; for example, the 32-story Rubber bearings have also been shown to soften in the vertical
Los Angeles City Hall, the 18-story Oakland City Hall, and nu- direction at large lateral deformations. In pure tension, rubber
merous projects in Japan. Certainly, overturning in these slender bearings tend to cavitate, or form small cavities in the rubber that
structures will generate large axial forces in the isolation bearings. blow out from negative pressure and link together to form cracks
Due to concerns about the stability of rubber isolation bearings in the rubber matrix. In one test, this occurred at negligible tensile
under large compressive loads and their ability to withstand ten- strains of about 0.0003, corresponding to tensile stress of only
sile loads, the peak axial forces in both tension and compression 1.59 MPa 共230 psi兲 共Clark et al. 1997兲. However, in recent
are of interest to the designer. projects, bearings under large lateral deformation were jacked up
Axial forces have been observed to significantly affect the re- 12– 20 mm 共0.5– 0.75 in.兲 with no evidence of cavitation damage
sponse of isolation bearings. First, a correlation between lateral 共Kelly 2003兲.
stiffness and axial load has been observed in several types of Although less documented, the yield strength of an LRB bear-
rubber bearings. In tests of disparate bearings—for example, ing, or strength of the lead core, has been observed to vary with
natural rubber, high-damping rubber 共HDR兲 and lead–rubber axial load, such that a lightly loaded bearing may not achieve its
共LRB兲 bearings—with both dowelled and bolted connections theoretical strength. This effect was first noted as an immediate
共Kelly et al. 1987; Griffith et al. 1988; Aiken et al. 1989兲, the decay in the lateral force–deformation hysteresis area in response
secant stiffness decreased with increasing axial load. As an ex- to a sudden drop in axial pressure 共Tyler and Robinson 1984兲. In
a three-story structure isolated with LRB bearings and subjected
1 to triaxial ground excitations, the bearings beneath heavily loaded
Assistant Professor, Dept. of Civil and Environmental Engineering,
4110 Old Main Hill, Utah State Univ., Logan, UT 84322-4110 interior columns showed much greater strength and energy dissi-
共corresponding author兲. E-mail: kryan@cc.usu.edu pation than identical bearings beneath lightly loaded exterior col-
2
Professor Emeritus, Pacific Earthquake Engineering Research Center umns 共Hwang and Hsu 2000兲.
共PEER兲, Univ. of California at Berkeley, 1301 S 46th St., Richmond, CA In spite of the preceding evidence, existing nonlinear bearing
94804-4698. E-mail: jmkelly@peer.berkeley.edu models do not account for the relation between axial loads and
3
Johnson Professor, Dept. of Civil and Environmental Engineering, lateral/vertical response of the bearings. The objective of this
Univ. of California at Berkeley, Berkeley, CA 94720-1710. E-mail: paper is to develop a model for LRB bearings that includes the
chopra@ce.berkeley.edu effect of axial loads, and a numerical implementation of this
Note. Associate Editor: Hayder A. Rasheed. Discussion open until model for dynamic analysis. The model is a nonlinear extension
May 1, 2006. Separate discussions must be submitted for individual pa-
of a two-spring model developed from linear stability theory of
pers. To extend the closing date by one month, a written request must be
filed with the ASCE Managing Editor. The manuscript for this paper was multilayer bearings 共Kelly 1997兲. Optionally included in the non-
submitted for review and possible publication on September 14, 2004; linearity is an empirical representation of the strength variation of
approved on January 26, 2005. This paper is part of the Journal of the lead core with axial loads. All behaviors of the model, which
Engineering Mechanics, Vol. 131, No. 12, December 1, 2005. ©ASCE, include variation of lateral stiffness and yield strength with axial
ISSN 0733-9399/2005/12-1270–1278/$25.00. load, and variation of vertical stiffness with lateral deformation,

1270 / JOURNAL OF ENGINEERING MECHANICS © ASCE / DECEMBER 2005

J. Eng. Mech. 2005.131:1270-1278.


Fig. 1. Response of axially loaded multilayer rubber bearing under
shear deformation, including tilting of rubber layers, in response to:
共a兲 compression loading and 共b兲 tensile loading. Note that middle
reinforcing layers tilt most.
Fig. 2. Two-spring model of isolation bearing in undeformed and
deformed configuration 关Reproduced from Kelly 共1997兲, Fig. 8-6,
Downloaded from ascelibrary.org by Iowa State University on 09/27/13. Copyright ASCE. For personal use only; all rights reserved.

with kind permission of Springer Science and Business Media兴


are confirmed by experimental data. From hereafter these behav-
iors are referred to as axial-load effects. Following development
of the algorithm to implement this bearing model, the response of the nominal vertical stiffness of the bearing, i.e., stiffness of the
the new model to a simple seismic pulse is explored. The model is additional vertical spring, is
less suitable for HDR bearings, because it lacks hardening at large
shear strains and other complex behavior typical of such bearings. E cA E cA s
kbzo = = 共2兲
tr hb

Stability Analysis of Multilayer Bearings where Ec = instantaneous compression modulus of the rubber–
steel composite bearing. In this context, the term nominal 共de-
In the stability analysis of multilayer bearings 共Koh and Kelly noted by subscript o兲 means absent axial-load effects.
1987; Kelly 1997兲, the bearing is treated as a continuous compos- If the shear stiffness of the two-spring model 共Fig. 2兲
ite system in which the steel layers do not deform, allowing pre- were infinite, the rotational stiffness divided by hb would
diction of the buckling load and the effective lateral stiffness in equal the conventional Euler buckling load PE = 共␲2 / h2b兲EIs. Here
the presence of axial load. The stability theory resembles the lin- EIs⫽ bending stiffness of a multilayer bearing 共Kelly 1997兲
earized theory of an elastic column, but accounts for shear defor-
1 hb
mation by considering rotation of the cross section, which is in- EIs = EcI 共3兲
dependent of the lateral deflection 共Koh and Kelly 1987兲. Also 3 tr
predicted by stability analysis 共Koh and Kelly 1987兲, the where I = conventional moment of inertia: ␲D4 / 64 or Ar2b;
multilayer bearing under simultaneous lateral and axial loading and rb = D / 4 = bending radius of gyration in terms of the bearing
undergoes an additional vertical displacement beyond that due to diameter D. Thus, the rotational stiffness, divided equally among
material axial flexibility. Demonstrated visually in Fig. 1, the ad- top and bottom springs 共Fig. 2兲, is PEhb.
ditional displacement ␦bz, either compressive or tensile depending With these simple linear constitutive relations, the equilibrium
on the corresponding axial load, is due to tilting of the middle equations relating the lateral force f b and axial 共compressive兲
bearing layers when the bearing is deformed in shear. force P to the deformation s across the shear spring and the ro-
tation ␪ through the rotational spring 共Fig. 2兲 are
Approximate Force–Deformation Relation Based
f b − kbos + P␪ = 0 共4a兲
on Linear Two-Spring Model
The preceding effects predicted by stability analysis can be rep- f bhb − PEhb␪ + P共s + hb␪兲 = 0 共4b兲
resented, approximately, by a simplified two-spring model of the
bearing 共Kelly 1997兲 that results in explicit force–deformation which assume small rotation ␪. The axial force P and the defor-
relations. The two-spring model 共Fig. 2兲 is a composition of rigid mation v across the vertical spring 共not shown in Fig. 2兲 are
tees connected by a rotational spring, subdivided at top and bot- related by an additional equation
tom, and a shear spring with frictionless rollers at midheight. The P − kbzov = 0 共4c兲
bottom plate is fixed and the top plate is constrained against ro-
tation. Axial flexibility of the bearing is included by an additional The kinematic equations relating the total lateral deformation
vertical spring in series 共not shown in Fig. 2兲. ub and vertical deformation ubz to the internal deformations s, ␪,
Assuming linear material behavior, the nominal shear stiffness and v, again assuming ␪ to be small, are
of a multilayer bearing, also the stiffness of the shear spring in
Fig. 2, is u b = s + h b␪ 共5a兲

GA GAs hb 2
kbo = = 共1兲 ubz = v + ␦bz = v + s␪ + ␪ 共5b兲
tr hb 2
where G = shear modulus; A = cross-sectional area; and tr = sum In Eq. 共5b兲, ubz, positive in compression, is the sum of v, the
thickness of the rubber layers. In some cases, it is convenient to deformation resulting from axial flexibility of the bearing, and
use modified area As = A共hb / tr兲, where hb⫽total height of the bear- ␦bz, the additional vertical displacement that occurs in the later-
ing, which accounts for the undeforming steel layers. Similarly, ally deformed configuration as shown in Figs. 1 and 2.

JOURNAL OF ENGINEERING MECHANICS © ASCE / DECEMBER 2005 / 1271

J. Eng. Mech. 2005.131:1270-1278.


Analysis of the linear two-spring model 共Kelly 1997兲 is sum-
marized in the following steps:
1. The critical buckling load is determined by setting the matrix
determinant of the system of equations in s and ␪ 关Eqs. 共4a兲
and 共4b兲兴 to zero

Pcr ⬇ ± 冑PS PE 共6兲


where PS = GAs = kbohb. Using the realistic assumption that
PE Ⰷ PS, Eq. 共6兲 is an approximation to the buckling load
determined from stability analysis of the multilayer bearing
关Kelly 共1997兲 关Eq. 共8.12兲兴兴.
2. The values of the shear deformation s and rotation ␪, found Fig. 3. 共a兲 Lateral force–deformation as function of P / Pcr and
by solving this system of equations 关Eqs. 共4a兲 and 共4b兲兴 are 共b兲 vertical force–deformation as function of lateral deformation ub.
Downloaded from ascelibrary.org by Iowa State University on 09/27/13. Copyright ASCE. For personal use only; all rights reserved.

substituted into Eq. 共5a兲, which, after approximation, simpli- Lateral and axial forces are normalized by static, or gravity load Pst.
fies to the following lateral force–deformation relation:

f b = k bu b where 冋 冉 冊册
kb = kbo 1 −
P
Pcr
2
共7兲
Written in flexibility form 关Eq. 共8兲兴, the vertical force–
deformation relation shows two distinct contributions to the ver-
tical deformation: a material effect and a second order geometric
Again, the lateral stiffness kb in Eq. 共7兲 for the two-spring effect. The additional vertical displacement ␦bz due to geometry
model is a good approximation to the stiffness derived from of deformation 关second term of Eq. 共8兲兴 increases according to the
stability analysis of the multilayer bearing 关Kelly 共1997兲, square of the lateral deformation. Note that even with no axial
Fig. 8-4兴. load on the bearing, the additional displacement ␦bz is nonzero,
3. These values of s and ␪, as well as the axial deformation v albeit small because PE Ⰷ PS, as shown in Fig. 3共b兲 at P = 0 for
determined from Eq. 共4c兲, are substituted into Eq. 共5b兲. This different lateral deformations ub. The net effect of this additional
leads to the following flexibility equation for vertical defor- displacement is an overall softening of the bearing in the vertical
mation as a function of force: direction, which depends on the lateral deformation 关Eq. 共10兲兴.
P 共PS + P兲 u2b This softening is also evident in the vertical force–deformation as
ubz = + 共8兲 a function of lateral deformation ub 关Fig. 3共b兲兴, wherein the slope
kbzo PE hb of the line for ub = 0 represents the nominal vertical stiffness 关Eq.
which is inverted to give 共2兲兴. Physically, the softening occurs as a result of the tilting of
bearing reinforcing layers 共Fig. 1兲, meaning the axial loads are
P = kbz ubz −冉 PS u2b
PE hb
冊 共9兲
resisted in part by shear.

with incremental, or tangent vertical stiffness kbz given by Variation of Yield Strength with Axial Load

kbz = 冉 1
+
kbzo PEhb
u2b
冊 −1

= kbzo 1 +
␲2r2b

3u2b −1
共10兲 Experimental evidence 共Tyler and Robinson 1984; Hwang and
Hsu 2000兲 of the bearing failing to achieve its full strength when
reduced to its final form by substituting Eq. 共3兲 for PE. Thus, lightly loaded has been attributed to lack of confinement of the
the lateral and vertical force–deformation relations of the lead plug. Skinner et al. 共1993兲 writes “the nominal upper limit of
bearing are defined by the coupled Eqs. 共7兲 and 共9兲. If axial- hysteretic force … should be achieved if there is no vertical slip-
load effects are neglected 关i.e., axial load P = 0 in Eq. 共7兲 and page of the plug sides and no horizontal slippage of the plug
lateral deformation ub = 0 in Eq. 共9兲兴, these equations reduce ends.” Current methods known to achieve the best confinement
to the familiar uncoupled linear force–deformation equations include minimizing the thickness of the rubber layers, fitting an
with nominal stiffnesses kbo 关Eq. 共1兲兴 and kbzo 关Eq. 共2兲兴. oversized lead plug into the undeformed cavity, and using a spe-
These stiffnesses are slightly modified in Ryan and Chopra cial technique to cap the plug. However, even recently tested
共2005兲 to include the effect of bulk compressibility. lead–rubber bearings seem to have this deficiency 共Hwang and
Lateral force–deformation curves for different axial loads are Hsu 2000兲.
shown in Fig. 3共a兲, with the curve corresponding to no axial load Because the preceding observation has not been justified by
共P = 0兲 representing the nominal stiffness. Each curve is linear mechanical theory, we have developed an empirical equation for
because P is constant. Together, the set of curves indicate the the yield strength as a function of the compressive load P based
reduction in lateral stiffness as the axial load increases. The stiff- on experimental data:
ness variation is inconsequential for typical design values of
Q = Qo共1 − e共−P/Po兲兲 共11兲
P / Pcr ⬍ 0.2, but the stiffness rapidly approaches zero as the axial
load approaches the critical load. where Qo = nominal yield strength of the bearing, achievable with
Consideration of the negative solution to Pcr 关Eq. 共6兲兴 leads to an adequate confining pressure; and the axial load Po, correspond-
a critical load in tension, hence the unconventional concept of ing to about 63% of nominal strength, should be chosen to match
tension buckling 共Kelly 2003兲. Because the two solutions to Pcr characteristic test data for the bearings. In Fig. 4, a plot of Eq.
are equal and of opposite sign, the bearing buckling behavior in 共11兲 shows that the strength declines quite rapidly for loads P
tension mirrors that in compression. Thus, the stiffness reduction below Po. The nominal strength is only achieved 共at 95% or
of Eq. 共7兲 is the same for both compressive and tensile loads greater level兲 when P 艌 3Po, therefore we recommend a minimum
共positive and negative values of P兲. of 3Po for the bearing design load Pst. When the bearing is in

1272 / JOURNAL OF ENGINEERING MECHANICS © ASCE / DECEMBER 2005

J. Eng. Mech. 2005.131:1270-1278.


Fig. 6. Vertical stiffness ratio kbz / kbzo versus deformation
Fig. 4. Influence of axial load P on yield strength Q 共normalized by bending radius兲 ub / rb for 共a兲 high damping rubber
Downloaded from ascelibrary.org by Iowa State University on 09/27/13. Copyright ASCE. For personal use only; all rights reserved.

number 1, 共b兲 high damping rubber number 2, and 共c兲 lead–rubber


bearings. Experimentally observed values for two different bearings
and two different load sequences are compared with Eq. 共10兲.
tension 共P ⬍ 0兲, the lead core is assumed to provide no strength,
and the effective yield strength is taken to be zero.
The proposed empirical model for strength variation with axial
load in LRB bearings 关Eq. 共11兲兴 is based on data for the LRB
Verification of Theory by Experimental Data bearing. To show that this model is justified by the data, the
observed strength ratio Q / Qo as a function of P / Po is compared
Experimental data that support the preceding theory are presented to Eq. 共11兲 for two LRB bearings 关Fig. 5共d兲兴. The nominal
briefly; the details are elaborated on elsewhere 共Ryan et al. 2004, strength Qo at each strain level is the observed strength of the first
Ryan and Chopra 2005兲. Fig. 5 compares the theoretical lateral bearing at its largest applied load 共3Pst = 235 kN兲, and the load Po
stiffness 关Eq. 共7兲兴 to the normalized postyield stiffness kb / kbo ob- for each bearing was selected visually to best fit the data. The
served in characteristic tests for three different bearings. Experi- proposed empirical model 关Eq. 共11兲兴 matches the bearing data
mental data for HDR bearings 关Figs. 5共a and b兲兴 match the theo- 关Fig. 5共d兲兴, especially at the larger strains where the data are most
retical stiffness 关Eq. 共7兲兴, especially at larger strains where the reliable. While the observed strength does not reduce to zero, it
data are considered to be most reliable. Data for the LRB bearing does fall exponentially with removal of axial load. It was docu-
关Fig. 5共c兲兴 does not support Eq. 共7兲 as well, partly due to the mented during testing that at zero axial load the lead core began
difficulty of obtaining a reliable estimate of the nominal stiffness to extrude out of both ends of the bearing, supporting our claim
kbo from the test data 共Ryan et al. 2004兲; however, the slope of the that the lead core is ineffective in tension. However, these LRB
experimental curves appears to match the slope of Eq. 共7兲. Con- bearings could not be tested in tension because they were attached
sidering the possible sources of experimental error, the theoretical by dowelled connections, which probably contributed to the ex-
model and experimental results seem to be in satisfactory agree- trusion of the lead core.
ment. Finally, Fig. 6 compares the vertical stiffness kbz / kbzo observed
at different imposed lateral strains to the theoretical vertical stiff-
ness as a function of lateral deformation 关Eq. 共10兲兴. The theoret-
ical and experimental stiffness ratios are in good agreement for
two of three bearings 关Figs. 6共a and c兲兴; we currently do not have
an explanation for the discrepancy in the observed stiffness of the
second bearing relative to the theoretical value 关Fig. 6共b兲兴. How-
ever, evaluated as a whole, the experimental data provide some
confirmation of the theory.

Nonlinear Extension of Two-Spring Model

The linear two-spring model adequately accounts for axial-load


effects in a bearing with linear material behavior. However, the
behavior of most isolation bearings is inherently nonlinear, espe-
cially LRB bearings due to the yielding of the lead core. This
requires extension of the two-spring model to include various
constitutive models for the shear spring. Three variations are de-
fined as follows:
1. Coupled linear model: the shear spring is linear; this model is
Fig. 5. Observed stiffness ratio kb / kbo versus P / Pcr for: 共a兲 high identical to the linear two-spring model 共Fig. 2兲;
damping rubber number 1, 共b兲 high damping rubber number 2, and 2. Coupled nonlinear constant-strength model: the shear spring
共c兲 lead–rubber bearings, compared with Eq. 共7兲. 共d兲 Observed follows a bilinear force–deformation relation; and
strength ratio Q / Qo versus P / Po for lead–rubber bearings compared 3. Coupled nonlinear variable-strength model: like the constant-
with Eq. 共11兲. Data points at P = 0, Pst / 2, Pst, 2Pst, 3Pst, and −Pst / 10, strength model, but the yield strength varies in time accord-
where Pst is design axial load. ing to Eq. 共11兲.

JOURNAL OF ENGINEERING MECHANICS © ASCE / DECEMBER 2005 / 1273

J. Eng. Mech. 2005.131:1270-1278.


To implement the strength variation due to axial load
共variable-strength model兲, the yield force f y in Eq. 共13b兲 is up-
dated consistent with the strength Q 关Eq. 共11兲兴 at each time in-
stant. The initial stiffness kI remains constant, determined by the
nominal strength Qo rather than the current strength Q. For both
the constant-strength and variable-strength models, f s共s兲 can be
computed to satisfy Eq. 共13兲 via the return mapping algorithm;
further details are given in Ryan and Chopra 共2005兲.

Numerical Implementation
An approach is presented to integrate the constant-strength and
Fig. 7. Lateral force–deformation as function of P / Pcr for: variable-strength bearing models into a typical dynamic analysis
共a兲 constant-strength and 共b兲 variable-strength models program. This approach is compatible with programs that com-
Downloaded from ascelibrary.org by Iowa State University on 09/27/13. Copyright ASCE. For personal use only; all rights reserved.

pute the deformations of the system at each time step, based on


The latter two models are hereafter abbreviated as the constant- the iterative stiffness method. Specifically, a local routine for the
strength and variable-strength models. Also defined for reference bearing is developed that, given the deformations, computes the
are comparable bearing models that neglect axial-load effects, bearing forces 关to satisfy the governing Eqs. 共12兲, 共4b兲, 共4c兲, and
such that the lateral force–deformation relation, either linear 共un- 共5兲兴 and the bearing stiffness matrix, which are returned for com-
coupled linear兲 or bilinear 共uncoupled nonlinear兲, is uncoupled putations at the system level.
from the vertical force–deformation relation, assumed to be lin-
Bearing Forces
ear. Thus, bilinear force–deformation 关see Fig. 7共a兲 for P = 0兴 is
Let F = 具f b , P典T represent the vector of independent bearing forces
used both for the shear spring in the constant-strength and
variable-strength models and as the complete lateral force– that are to be computed in the local bearing routine. The resultant
deformation in the uncoupled nonlinear model. moment M 1, found by equilibrium of the bearing in the deformed
configuration, should also be applied at the top of the bearing

Governing Equations M 1 = 共f bhb + Pub兲 共14兲


For the coupled linear model, the governing equations 关Eqs. 共4兲 The governing equations 关Eqs. 共12兲, 共4b兲, 共4c兲, 共5a兲, and 共5b兲兴
and 共5兲兴 are unchanged. Note that we do not use the approxima- represent a system of five nonlinear equations in five unknowns:
tions that led to Eqs. 共7兲 and 共9兲. To consider a different consti- f b, P, s, ␪, and v. This system of equations is recast in root-
tutive model for the shear spring, Eq. 共4a兲 becomes finding form

冦冧冦 冧
f b − f s共s兲 + P␪ = 0 共12兲 g1 f b − f s共s兲 + P␪
where the linear shear spring has been replaced by a general force g2 f bhb − PEhb␪ + P共s + hb␪兲
f s共s兲.
g= g3 = P − kbzov 共15兲
The force of the shear spring in the constant-strength model
may be determined numerically by classical rate-independent uni- g4 u b − s − h b␪
directional plasticity 共Simo and Hughes 1998兲, where the force– g5 ubz − s␪ − hb␪2/2 − v
deformation relation is elastic–plastic with kinematic hardening,
and solved by Newton’s method, i.e., find x = 具f b , P , s , ␪ , v典T to
and thus governed by the following constitutive law, yield func-
satisfy g共x兲 = 0.
tion, flow rule, and hardening law:
The converged solution x at the previous global iteration
f s = kI共s − s p兲 共13a兲 serves as an initial guess x共0兲. An improved solution is found by

⌽共f s,q兲 = 兩f s − q兩 − f y 艋 0 共13b兲 x共k兲 = x共k−1兲 − 关J共x共k−1兲兲兴−1g共x共k−1兲兲 共16兲


for k = 1 , 2 , . . ., where the Jacobian J共x兲 共Jij = ⳵gi / ⳵x j兲, determined
ṡ p = ␥˙ sgn共f s − q兲 共13c兲 from Eq. 共15兲

冤 冥
q̇ = ␥˙ H sgn共f s − q兲 共13d兲 ⳵fs
1 ␪ − P 0
⳵s
respectively. The constitutive law 关Eq. 共13a兲兴 shows that the
spring force equals the initial stiffness kI times the elastic compo- hb 共s + hb␪兲 P 共P − PE兲hb 0
J共x兲 = 共17兲
nent of deformation s − s p, s p being the plastic deformation. The 0 1 0 0 − kbzo
initial stiffness kI = kbo + Qo / sy depends on the nominal stiffness 0 0 −1 −h 0
and strength, as well as the yield deformation sy. The yield func-
0 0 − ␪ − 共s + hb␪兲 − 1
tion ⌽ 关Eq. 共13b兲兴 determines the set of admissible forces, where
the back force q stores the translation of the yield surface, and the has a nonzero determinant, and hence is invertible, even when P,
yield force f y = Qo + kbosy. The spring response is elastic inside the s, and ␪ are zero. Eq. 共16兲 is applied repeatedly until the incre-
yield surface 共⌽ ⬍ 0兲, and plastic flow occurs on the yield surface mental change in the solution is less than a desired tolerance, that
共⌽ = 0兲, determined by the flow rule 关Eq. 共13c兲兴 with constant slip is: 储x共k兲 − x共k−1兲储 ⬍ tol. At each iteration the shear spring force f s共s兲
rate ␥˙ . Evolution of the back force is governed by the hardening in Eq. 共15兲 and tangent ⳵ f s / ⳵s in Eq. 共17兲 are computed by the
law 关Eq. 共13d兲兴, with hardening stiffness H = kIkbo / 共kI − kbo兲. return mapping algorithm 共Ryan and Chopra 2005兲.

1274 / JOURNAL OF ENGINEERING MECHANICS © ASCE / DECEMBER 2005

J. Eng. Mech. 2005.131:1270-1278.


Bearing Stiffness Matrix and
Next, the bearing stiffness matrix kb, relating a change in the
bearing forces dF= 具df b , dP典T to the change in deformations

冤 冥
1 ␪
dU = 具dub , dubz典T, is derived in three steps:
1. Take differentials of the equilibrium equations 关Eqs. 共12兲, T = hb 共s + hb␪兲 共19b兲
共4b兲, and 共4c兲兴, resulting in 0 1
2. Take differentials of the kinematic equations 关Eq. 共5兲兴, result-
keqdv = TdF 共18兲 ing in

dU = TTdv 共20兲
where dv = 具ds , d␪ , dv典T and the matrices keq and T are given
by 3. Substitute dv from Eq. 共18兲 into Eq. 共20兲
Downloaded from ascelibrary.org by Iowa State University on 09/27/13. Copyright ASCE. For personal use only; all rights reserved.

冤 冥
⳵fs dU = 共TTfeqT兲dF 共21兲
−P 0
⳵s −1
keq = 共19a兲 where feq = keq . The resultant flexibility matrix fb = TTfeqT,
− P 共PE − P兲hb 0 which relates the incremental deformations dU to the incre-
0 0 kbzo mental forces dF, is given explicitly as

冤 冥
共PE + P兲hb + h2b共⳵ f s/⳵s兲 PEhb␪ + 关P + hb共⳵ f s/⳵s兲兴共s + hb␪兲
hb共⳵ f s/⳵s兲共PE − P兲 − P 2
hb共⳵ f s/⳵s兲共PE − P兲 − P2
fb = 共22兲
PEhb␪ + 共P + hb⳵ f s/⳵s兲共s + hb␪兲 共PE + P兲hb␪2 + 2Ps␪ + 共⳵ f s/⳵s兲共s + hb␪兲2 1
+
hb共⳵ f s/⳵s兲共PE − P兲 − P2 hb共⳵ f s/⳵s兲共PE − P兲 − P2 kbzo

Given the complexity of Eq. 共22兲, the bearing stiffness matrix


kb, the inverse of fb, is not derived explicitly. However, the fol-
lowing observations are relevant:
k11 ⬇
共PS PE − P2兲
P Eh b
冋 冉 冊册
= kbo 1 −
P
Pcr
2
共25兲

1. Suppose both the axial load P and deformations s and ␪ are which is identical to the approximate lateral stiffness kb in
zero, and the shear spring is linear 共⳵ f s / ⳵s = Ps / hb兲; then the Eq. 共7兲.
flexibility matrix 关Eq. 共22兲兴 is diagonal with elements 3. Suppose hb␪ is small relative to the total lateral deformation
ub, the shear spring is linear, and P Ⰶ Pcr; then f 22 can be
共PE + PS兲hb approximated as
f 11 = 共23a兲
PE PS
1 u2b 1
and f 22 ⬇ + 共26兲
PE hb kbzo

1 whose inverse is the tangent vertical stiffness kbz derived


f 22 = 共23b兲 earlier 关Eq. 共10兲兴.
kbzo
Thus, the exact stiffness matrix kb has been shown with relevant
Typically PE Ⰷ PS, giving f 11 ⬇ hb / PS = 1 / kbo, such that the assumptions to lead to the same approximate force–deformation
lateral and vertical bearing stiffnesses reduce to their nomi- equations 关Eqs. 共7兲 and 共9兲兴 given earlier, a good check on its
nal values 共k11 ⬇ kbo 关Eq. 共1兲兴 and k22 = kbzo 关Eq. 共2兲兴兲. accuracy. An outline of the complete bearing routine, including
2. Suppose only the deformations s and ␪ are zero, and the solving for the force vector F and stiffness matrix kb, is given in
shear spring is linear; then the flexibility matrix is diagonal Ryan and Chopra 共2005兲. In addition, the bearing routine is ex-
with pected to be implemented and accessible as a material model in
the open source structural and earthquake analysis program Op-
共PE + P + PS兲hb enSees 共McKenna and Fenves 2001兲 at the time of press.
f 11 = 共24a兲
PS PE − PS P − P2
and Observed Response of New Bearing Models

1 Lateral and Vertical Force–Deformation Trends


f 22 = 共24b兲
kbzo
Lateral force–deformation behavior for the coupled linear model
Inverting f 11 and again taking PE Ⰷ PS gives was already demonstrated in Fig. 3共a兲, where the lateral stiffness

JOURNAL OF ENGINEERING MECHANICS © ASCE / DECEMBER 2005 / 1275

J. Eng. Mech. 2005.131:1270-1278.


Fig. 8. 共a兲 Acceleration ügo, 共b兲 velocity u̇go, and 共c兲 displacement
ugo history of seismic excitation: approximately 1.5 cycles of
sinusoidal pulse with velocity u̇go = 90 cm/ s and period T p = 4 s.
Downloaded from ascelibrary.org by Iowa State University on 09/27/13. Copyright ASCE. For personal use only; all rights reserved.

Isolation period Tb for system excited is 2 s.

was shown to decrease as the applied force P approached Pcr. The


lateral force–deformation relation of the constant-strength model
关Fig. 7共a兲兴 is closely related to the coupled linear model. The
postyield stiffness in Fig. 7共a兲 is essentially identical to the stiff- Fig. 10. Response of left exterior bearing—using constant-strength
ness of the coupled linear model 关Fig. 3共a兲兴, showing the same model—of rigid block subjected to seismic pulse: 共a兲 lateral and
successive decline as the axial force is increased toward Pcr. The vertical deformation and force histories 共qualitative, with relative
initial stiffness and energy dissipated in a single cycle 共loop area兲 scale indicated兲, 共b兲 lateral force–deformation, and 共c兲 vertical
are affected by axial force only negligibly. As a special case, the force–deformation. Also shown is comparative response with
curve with P = 0 is equivalent to the uncoupled nonlinear model, uncoupled nonlinear model; Pst / Pcr = 0.4.
which neglects axial-load effects.
The primary effect of including the strength variation of Eq.
共11兲 共variable-strength model兲 is that the yield strength, and hence
bearing should achieve a balance, in the sense that P should be
energy dissipated in a single cycle, also depend on the axial force
large enough to avoid significant strength degradation but well
P 关Fig. 7共b兲兴. Recall that when P = 0 or small, the strength, and
below Pcr.
hence energy dissipated, is also zero or small, as observed in
For an imposed lateral deformation, the vertical force–
Fig. 7共b兲. As the axial force becomes large relative to Po and Pcr,
deformation relation is independent of the particular bearing
the strength is regained but the postyield stiffness decreases simi-
model 共linear, constant strength or variable strength兲, and the
lar to the constant-strength model 关Fig. 7共a兲兴. A well-designed
curves of Fig. 3共b兲 for the coupled linear model also represent the
constant-strength and variable-strength models. Clearly, the verti-
cal force–deformation will vary between these curves when the
lateral deformation varies freely during seismic loading.

Response to Seismic Pulse


For each variation of the bearing model that has been presented,
a rigid block supported on two bearings is subjected to the seis-
mic pulse of Fig. 8, and the response of the left exterior bearing
is shown 共Figs. 9–11兲. These figures demonstrate response histo-
ries of the lateral and vertical deformations and forces, the lateral
force–deformation relation, and the vertical force–deformation
relation. Lateral and vertical forces are normalized by the static
load Pst.
The bearing response predicted by the coupled linear model is
given 共Fig. 9兲, with the comparative response using the uncoupled
linear model shown for reference. The lateral deformation and
axial force of the bearing are nearly in phase 关Fig. 9共a兲兴. As a
consequence, the axial force P increases when the lateral defor-
mation is large and positive, and an associated drop in stiffness
关Eq. 共7兲兴 is observed in the lateral force–deformation relation
关Fig. 9共b兲兴. The lateral stiffness does not appear to change at
Fig. 9. Response of left exterior bearing—using coupled linear negative lateral deformations, because the axial force tends to-
model—of rigid block subjected to seismic pulse: 共a兲 lateral and ward zero such that the lateral stiffness defaults to its nominal
vertical deformation and force histories 共qualitative, with relative value. Including axial-load effects 共coupled linear versus un-
scale indicated兲, 共b兲 lateral force–deformation, and 共c兲 vertical coupled linear model兲 appears to increase the peak lateral defor-
force–deformation. Also shown is comparative response with mation by about 20%.
uncoupled linear model; Pst / Pcr = 0.4. The vertical deformation is considerably greater when axial-

1276 / JOURNAL OF ENGINEERING MECHANICS © ASCE / DECEMBER 2005

J. Eng. Mech. 2005.131:1270-1278.


disordered, it appears hysteretic in nature relative to the coupled
linear model. Perhaps since the axial force variation resembles the
lateral force variation, the vertical hysteresis is related to hyster-
etic behavior in the lateral direction 关Fig. 10共b兲兴. Large local
variations in the vertical force–deformation relation are attributed
to the high frequency component of axial force. The peak vertical
deformation is significantly less than that of the coupled linear
model 关Fig. 9共c兲兴 due to the reduction of peak lateral deformation.
The bearing response predicted by the variable-strength
model, including response histories 关Fig. 11共a兲兴 and force–
deformation relations 关Figs. 11共b and c兲兴—which are compared
with the response predicted by the uncoupled nonlinear
model—is similar to that of the constant-strength model 共Fig. 10兲.
The influence of variable strength 关Eq. 共11兲兴 is readily apparent
Downloaded from ascelibrary.org by Iowa State University on 09/27/13. Copyright ASCE. For personal use only; all rights reserved.

when the axial force is close to zero by comparing Figs. 10共b兲 and
11共b兲. Observe that the strength decreases and the yield surface
contracts when the lateral force is negative 关Fig. 11共b兲兴. Com-
pared to the uncoupled nonlinear model, the average width of the
force–deformation loop is smaller, and the associated decrease in
energy dissipation causes the peak lateral deformation to increase,
Fig. 11. Response of left exterior bearing—using variable-strength which in turn increases the peak vertical deformation 关Fig. 11共c兲兴.
model—of rigid block subjected to seismic pulse: 共a兲 lateral and
The local oscillations in the lateral force–deformation are stronger
vertical deformation and force histories 共qualitative, with relative
compared to Fig. 10共b兲 as the behavior is affected by local
scale indicated兲, 共b兲 lateral force–deformation, and 共c兲 vertical
strength variations as well as stiffness variations.
force–deformation. Also shown is comparative response with
uncoupled nonlinear model; Pst / Pcr = 0.4 and Pst / Po = 3.

Conclusions
load effects are included 关coupled linear model versus uncoupled
linear model, Figs. 9共a and c兲兴. To understand this, recall that the This investigation to develop a nonlinear model for isolation bear-
vertical force–deformation relation depends greatly on the lateral ings that includes the influence of axial load has led to the fol-
deformation 关Eq. 共9兲兴. Thus, when the axial force is close to lowing conclusions:
its static value, the lateral deformation is close to zero 共recalling 1. From testing of both high-damping rubber and lead–rubber
the correlation between P and ub observed above兲 and the vertical bearings, it has been observed that the lateral stiffness de-
force–deformation is essentially that of the uncoupled linear creases with increasing axial load, the lateral yield strength
model. However, as the axial force deviates from the static decreases with decreasing axial load 共lead–rubber bearings
force in either direction, the lateral deformation becomes large, only兲, and the vertical stiffness decreases with increasing lat-
resulting in vertical softening and a considerable increase in ver- eral deformation.
tical deformation, and thus the arc-shaped vertical force– 2. Based on linear stability theory of an elastic column, a two-
deformation 关Fig. 9共c兲兴, with vertical deformation increasing at spring model of the bearing, consisting of a shear spring and
axial forces larger or smaller than Pst. Due to the inherently linear a rotational spring divided at top and bottom 共Fig. 2兲, accu-
relation between axial force P and deformation ubz 关Eq. 共9兲兴, rately accounts for axial-load effects in linear models of
the total deformation ubz is obviously greater when P ⬎ Pst than bearings. However, bearings with high-damping fillers or
when P ⬍ Pst. lead cores that provide energy dissipation are nonlinear and
The bearing response predicted by the constant-strength model should be modeled as such.
is plotted in Fig. 10 and compared with the response predicted by 3. Such modeling is achieved by extending the two-spring
the uncoupled nonlinear model. Whereas for the linear models the model to include a nonlinear constitutive model for the shear
axial force history and lateral deformation history were in phase spring. To form the constant-strength bearing model, this
关Fig. 9共a兲兴, for these nonlinear models the axial force history is shear spring was modeled by unidirectional plasticity with
more closely in phase with the lateral force history 关Fig. 10共a兲兴. kinematic hardening. Numerically, the model is implemented
For the constant-strength model, this causes the lateral stiffness by viewing the equilibrium 关Eqs. 共12兲, 共4b兲, and 共4c兲兴 and
共initial or postyield兲 to show the greatest decrease at large positive kinematic 关Eq. 共5兲兴 equations as a system of five nonlinear
lateral forces 关Fig. 10共b兲兴, which correspond to the maximum equations, which are solved by Newton’s method for the cur-
axial forces. The result is a slight increase in peak lateral defor- rent forces in the bearing. The instantaneous stiffness matrix
mation compared to the uncoupled nonlinear model. Because of is derived by taking differentials of these equations. Al-
the changing stiffness, the apparent width of the hysteresis loop though an improvement, the model is an incomplete repre-
changes during the deformation cycle, whereas this width is con- sentation of high-damping rubber bearings because it ne-
stant for the uncoupled nonlinear model. A high-frequency com- glects complex effects like strain hardening.
ponent is observed in the axial force cycle 关Fig. 10共a兲兴, causing 4. An empirical model 关Eq. 共11兲兴 was developed to account for
local stiffness variations and small oscillations in the lateral the varying yield strength in lead–rubber bearings, which can
force–deformation relation 关Fig. 10共b兲兴. be calibrated to match the experimentally observed bearing
The vertical force–deformation relation of the constant- response. This behavior is optionally included by simply up-
strength model is difficult to interpret 关Fig. 10共c兲兴. Although quite dating the yield force in the plasticity equations to reflect the

JOURNAL OF ENGINEERING MECHANICS © ASCE / DECEMBER 2005 / 1277

J. Eng. Mech. 2005.131:1270-1278.


time-varying yield strength, leading to the variable-strength Hwang, J.-S., and Hsu, T.-Y. 共2000兲. “Experimental study of isolated
model. building under triaxial ground excitations.” J. Struct. Eng., 126共8兲,
5. The constant-strength and variable-strength models were 879–886.
demonstrated in the response of a rigid block supported Kelly, J. M. 共1997兲. Earthquake-resistant design with rubber, 2nd Ed.,
by two bearings subjected to a seismic pulse. The effects Springer-Verlag, London.
were most pronounced in the variable-strength model, where Kelly, J. M. 共2003兲. “Tension buckling in multilayer elastomeric bear-
the peak lateral deformation increased by about 20% and ings.” J. Eng. Mech., 129共12兲, 1363–1368.
Kelly, J. M., Buckle, I. G., and Koh, C.-G. 共1987兲. “Mechanical charac-
the peak vertical deformation more than doubled compared
teristics of base isolation bearings for a bridge deck model test.” Rep.
to the uncoupled nonlinear model 共i.e., without axial-load
No. UCB/EERC-86/11, Earthquake Engineering Research Center,
effects兲. Univ. of California, Berkeley, Calif.
Koh, C.-G., and Kelly, J. M. 共1987兲. “Effects of axial load on elastomeric
isolation bearings.” Rep. No. UCB/EERC-86/12, Earthquake Engi-
Acknowledgment neering Research Center, Univ. of California, Berkeley, Calif.
McKenna, F., and Fenves, G. L. 共2001兲. OpenSees, the Open System for
Downloaded from ascelibrary.org by Iowa State University on 09/27/13. Copyright ASCE. For personal use only; all rights reserved.

The writers are grateful for the funding of this research contrib- Earthquake Engineering Simulation 具http://opensees.berkeley.edu,
uted by a California state grant. 2004典.
Ryan, K. L., and Chopra, A. K. 共2005兲. “Estimating the seismic response
of base-isolated buildings including torsion, rocking, and axial-load
References effects.” Rep. No. UCB/EERC-2005/01, Earthquake Engineering Re-
search Center, Univ. of California, Berkeley, Calif.
Aiken, I. D., Kelly, J. M., and Tajirian, F. F. 共1989兲. “Mechanics of low Ryan, K. L., Kelly, J. M., and Chopra, A. K. 共2004兲. “Experimental
shape factor elastomeric seismic isolation bearings.” Rep. No. UCB/ observation of axial-load effects in isolation bearings.” Proc., 13th
EERC-89/13, Earthquake Engineering Research Center, Univ. of Cali- World Conf. on Earthquake Engineering, Paper No. 1707, Canadian
fornia, Berkeley, Calif. Association for Earthquake Engineering, Vancouver, British Colum-
Clark, P. W., Aiken, I. D., and Kelly, J. M. 共1997兲. “Experimental studies bia, Canada.
of the ultimate behavior of seismically-isolated structures.” Rep. No. Simo, J. C., and Hughes, T. J. R. 共1998兲. Computational inelasticity,
UCB/EERC-97/18, Earthquake Engineering Research Center, Univ. of Springer, New York.
California, Berkeley, Calif. Skinner, R. I., Robinson, W. H., and McVerry, G. H. 共1993兲. An intro-
Griffith, M. C., Kelly, J. M., Coveney, V. A., and Koh, C.-G. 共1988兲. duction to seismic isolation, Wiley, New York.
“Experimental evaluation of medium-rise structures subject to uplift.” Tyler, R. G., and Robinson, W. H. 共1984兲. “High-strain tests on lead-
Rep. No. UCB/EERC-88/02, Earthquake Engineering Research Cen- rubber bearings for earthquake loadings.” Bull. New Zealand Nat.
ter, Univ. of California, Berkeley, Calif. Soc. Earthquake Eng., 17共2兲, 90–105.

1278 / JOURNAL OF ENGINEERING MECHANICS © ASCE / DECEMBER 2005

J. Eng. Mech. 2005.131:1270-1278.

You might also like