You are on page 1of 19

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE.

For personal use only; all rights reserved.

ELASTOMERIC BEARINGS: STATE-OF-THE-ART


By Charles W. Roeder 1 a n d J o h n F. Stanton, 2 M e m b e r s , ASCE
ABSTRACT: Elastomeric bearings are widely used by structural engineers, but
their behavior is not well understood. The paper provides a brief description
of the state of knowledge throughout the world with respect to these bearings.
It summarizes the material behavior of elastomers and the theoretical and experimental research on bearings. It notes the various modes of failure, and the
major design methods are described and compared. There are wide variations
in standard practice throughout the world, and the contradictions and reasons
behind the variations are noted. The objective of the paper is to help the structural engineer get a better understanding of the design and selection of elastomeric bearings.
INTRODUCTION

Elastomeric bearings are widely used in structures. In bridges they


accommodate movements such as creep and thermal expansion, and in
precast concrete construction they act as seating pads which provide
uniform bearing for members, absorbing small movements and fabrication misalinement. Further, they are being used increasingly for seismic base isolation and machine vibration control. These bearings are economical and maintenance free and they have developed a history of
satisfactory performance with few problems. Although their use has increased dramatically in recent years, their behavior is complex and is
not well understood by structural engineers, and thus bearings may not
live up to the designer's expectations.
There are several reasons for this lack of understanding. First, the material properties of elastomers are very different from those commonly
encountered by structural engineers. Second, the mechanics of elastomeric bearings are also unusual, because of the different configurations
employed and the large strains which may occur.
This paper attempts to resolve these misunderstandings. It describes
the behavior of elastomeric bearings and it presents a review of the
properties of the elastomer. Results of theoretical analyses and experimental studies are also presented. Potential failure modes, important
design criteria, design methods and specifications are described and
compared, and their rationale is examined. Finally, the limitations of
existing knowledge and research needs are noted. This paper will emphasize bridge bearings, but the findings are also applicable to other
structural uses of elastomeric bearings.
MATERIAL PROPERTIES

Elastomers are very different from other structural materials and some
of their characteristics are summarized in Table 1. Nearly all elastomeric
'Assoc.
Prof., Dept. of Civ. Engrg., Univ. of Washington, Seattle, Wash. 98195.
2
Asst. Prof., Dept. of Civ. Engrg., Univ. of Washington, Seattle, Wash. 98195.
Note.Discussion open until May 1,1984. 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 review and
possible publication on September 30, 1982. This paper is part of the Journal of
Structural Engineering, Vol. 109, No. 12, December, 1983. ASCE, ISSN 07339445/83/0012-2853/$01.00. Paper No. 18470.
2853

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

TABLE 1.Summary of Material Properties of Natural Rubber and Chloroprene


\

Property

Materials^.

(D

Low temperature behavior


(2)

Deterioration
due to environmental conditions

0)

Chloroprene

Less susceptible
Stiffens at
than natural
higher temrubber
peratures
than natural
rubber and
exhibits both
first and second order
transition

Natural rubber

Less susceptiSusceptible to
ble to low
ozone cracktemperatures.
ing but conSecond order
trolled by
transition
anti-ozonant.
only at apMore suscepprox -40 F
tible than
chloroprene

Creep and
stress
relaxation
(4)

Mechanical
properties
(5)

Rate dependent
behavior
(6)

High temperature
behavior
(7)

Both mate- Both maBoth mateNonlinear


rials
terials
rials
time and
stiffen
become
creep
temperaunder
more
similar
ture dedynamic
flexible
amounts.
pendent.
loads
at high
Typically
Properties
tempera25% or
vary with
tures
less for
compound
with inmost combut simipounds.
lar for
creasing
If anyboth
creep
thing
materials
and renatural
laxation
rubber
creeps
slightly
less.

bearings are made of either natural rubber or a synthetic rubber such as


chloroprene. Other rubbers have been tried (1,9,47), but only these two
have developed a history of satisfactory performance and are widely accepted by the profession. They have highly nonlinear, time and temperature dependent stress-strain behavior. Nonlinearity is introduced by
creep, stress relaxation and large strains, since tensile strains (60) (elongation at break, EB) of up to 600% are possible. Elastomers stiffen at low
temperatures (13,47,48) and under dynamic loading (59). Stiffening at
low temperatures is a complex phenomenon, but it is generally agreed
that natural rubber is less susceptible to it than chloroprene.
The raw rubber must be compounded to form the elastomer, which
must be vulcanized after the reinforcement has been bonded to the rubber. This assures a stable, durable bearing with a high quality bond between the elastomer and the reinforcement. During compounding, the
rubber is mixed with fillers such as carbon black, oils and other additives
which aid the manufacturing process, protective systems such as antiozonants and antioxidants, and a vulcanizing agent such as sulfur (28).
The mix is then vulcanized by the application of heat and pressure. The
details of specific elastomer compounds vary and are frequently regarded as proprietary information by the bearing manufacturer. However, they affect the strength, stiffness, and other material properties of
the rubber, and so the mechanical properties of the finished bearing will
vary significantly from manufacturer to manufacturer. This is one of the
major problems facing structural engineers in the design and selection
of elastomeric bearings. It can be primarily attributed to the fact that
structural engineers and bearing manufacturers have different interests
and speak different languages. The structural engineer is interested in
providing a bearing device which accommodates movements, and he is
concerned with the strength and stiffness of the bearing. The manufacturer produces the bearing to obtain specific characteristics of the rubber,
such as hardness and elongation at break. Thus, the structural engineer
2854

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

must specify these rubber characteristics even though he does not always understand their meaning or relationship to the structural' performance of the bearing.
While elastomers are highly nonlinear materials, the stress-strain relationships actually used in the design and analysis of bearings assume
isotropic linear elastic behavior. This assumption is clearly not correct,
but it is easy to use and provides adequate accuracy over the range of
application. For design purposes, an elastic modulus, , shear modulus,
G, and Poisson's ratio, v, must normally be defined. Elastomers are virtually incompressible but are flexible under uniaxial stress. Thus
v - 0.5
G

(1)

(2)

"i

and E = a small number [typically less than 1,000 psi (6.9 MPa)]. Variations from Eqs. 1-2 are sometimes used in practice, but these are empirical relations which account for factors other than true material properties. As noted earlier, the structural designer is interested in the stiffness
(i.e., E and G) of the elastomer, but he must also specify rubber properties such as the hardness and elongation at break for the manufacturer.
The hardness of the elastomer must normally be specified by the structural engineer. Both the Shore A Durometer and International Rubber
Hardness scales are used today, but they are nearly identical over the
range 50-60 which is most commonly used in elastomeric bearings. For
design purposes, the material moduli, E and G, are correlated to hardness. Generally, harder rubbers are stiffer with a reduced elongation at
break. Gent (21) used the theory of elasticity to develop the more precise
relationship between elastic modulus and hardness shown in Fig. 1.
However, for practical bearings it must be noted that variability in the
hardness measurement and correlation between hardness and the elastic
modulus indicates that E and G can only be approximately predicted for
a given elastomeric bearing.
GENERAL BEHAVIOR

The behavior of elastomeric bearings is dependent upon the material


properties of the rubber. Since these materials are stiff against volumetric changes but flexible in uniaxial compression, large strains with
very little change in volume are to be expected. Reinforced bearings (58)
have steel plates placed in the bearing in layers as shown in Fig. 2. The
reinforcement has large in-plane stiffness compared to the elastomer,
and, thus, the attached rubber can develop only out-of-plane shear strains.
This results in bulging of the elastomer as shown in Fig. 3(a) when the
bearing is compressed. Lateral translation of the bearing produces a simple shear deformation with little or no bulging as shown in Fig. 3(c),
and rotation results in bulging as shown in Fig. 3(b). The magnitude of
the bulging depends on the geometry of the bearing and the properties
of the elastomer. Stiff rubbers bulge less than flexible ones, and thick
layers bulge more than thin layers with the same plan dimensions. Since
increased bulging induces larger shear strains in the elastomer as shown
2855

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

100-

20
40 60
80 100
INTERNATIONAL RUBBER
HARDNESS (DEGREES)

FIG. 1.Elastic Modulus as Function


of Rubber Hardness

FIG. 2.Typical Elastomeric Bearing

in Fig. 3, it must be controlled. Increasing the stiffness with a harder


rubber reduces the shear strains, but it also makes the bearing less effective in accommodating structural movements. Thus, the ideal method
for controlling the behavior of an elastomeric bearing is to adjust the
geometry to the specific condition.
The shape factor, S, is used to model this geometric effect, where for
one layer of a reinforced bearing
S=

Loaded Area
Area Free to Bulge

(3)

Introduction of many reinforcement layers will produce bearings with


COMPRESSIVE LOAD

, i ,

.r*

0
ROTATION OR
^

APPLIED MOMENT

^=^
< ^

\=

SHEAR FORCE

FIG. 3.Typical Deformations and Shear Strains in Bearings

2856

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

large shape factors, reduced bulging and shear strains, increased stiffness in compression and rotation, and no significant change in the resistance to translational movement. The shear stress in the elastomer
produces tensile stress in the reinforcement, and both are proportional
to the mean compressive stress. Steel is strong in tension and, thus, few
problems have occurred due to tensile stress. Fiberglass fabric reinforcement has also been used in the United States. It is much weaker than
steel in tension, and thus usually controls the strength of the bearing.
Note that this review does not apply to pads.made of proprietory materials consisting of elastomer reinforced with randomly oriented fibers.
Unreinforced pads (58) are elastomeric bearings without internal reinforcement. These pads also bulge when loaded, but the bulge restraint
and shear strains are developed by friction between the bearing and the
load surface rather than by bond to reinforcement. Friction of rubber in
contact with steel or concrete is highly variable and is quite different
from that of other materials. Thus, slippage (2) at the load surface frequently occurs. Slippage results in a more flexible bearing and greatly
increased elastomer strains. Most design codes either do not permit (30)
or severely penalize (10,16,17) the use of unreinforced pads. The penalty
usually takes the form of decreasing the shape factor by a factor of 1.8
for unreinforced pads.
THEORETICAL DEVELOPMENTS

At least four (11,27,53,59) different linear elastic, isotropic, infinitesimal strain theories have been derived for predicting the behavior of elastomeric bearings under compressive load. These theories start with slightly
different assumptions but produce similar results and so only one will
be reviewed in this paper. In each, compressive deflection is largely accounted for by lateral bulging of the elastomer as a parabola. The result
(11) can be expressed as
(4)

fcE
and

*% = &o(5)
for a rectangular bearing in which <r, e, and TZXIMX = respectively, the
average compressive stress, average compressive strain, and maximum
shear stress. The dimensionless coefficients fc and gc are found by an
infinite series. For a rectangular bearing of length, L, width, W, and
layer thickness, t:
32 II

tanh 6,
l -

1
=,

(6)

(1 - sech e)

1T /f

4 \L
.i%RW

1 -

(7)

tanh 6,
2857

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

>-m' :
and

nRTtW
6 = 2L

(9)
'

Note that the bulk modulus, K, is infinite if v = 0.5. Fig. 4 shows plots
of typical values of fc and gc for rectangular bearings. Similar linear elastic
solutions have been derived for nonrectangular bearings in compression
(27) and bearings under rotation (11,27).
The previously mentioned theory is valid for very small deflections,
and Gent (20) developed a theory which approximately models nonlinear behavior with the bulk compression effect. Others (32,37,40) developed more formalized finite strain solutions. While the infinitesimal strain
theories are of limited applicability, they provide some very valuable
information. They show that large shear stress and strain are present
within the rubber and that large tensile stress (58) is present in the reinforcement. Further, these theories have served as the basis of numerous appropriate provisions in design codes (58).
It is generally agreed that the only practical model for predicting translational movement of the bearing as shown in Fig. 3(c) is the assumption
of simple shear, i.e.
Fs = GAy,

(10)

in which Fs = the transverse force on the bearing; A = the area of the


bearing; ys = the shear strain (i.e., 7S = As/T); and T = the total effective
rubber thickness. Simple shear is theoretically incorrect since it does not
satisfy equilibrium at the free edge where the shear stress must equal
zero, but experiments (26) suggest that the model is adequate for prac-

SHAPE FACTOR

SHAPE FACTOR

FIG. 4.(a) Dlmensionless Parameters as Function of Shape Factor; (b) Dimenslonless Parameters as Function of Shape Factor
2858

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

tical applications. However, the behavior under a combination of different types of loadings is much less clear. Superposition of solutions is
usual for solving combined load problems with linear elastic materials,
but superposition does not apply to elastomeric bearings since shear strains
in excess of 100% are probable under working conditions. The effect of
combining compression with shear has been investigated (51) and it has
been suggested that the material property, G, remains constant through
a wide range of strain and load conditions. However, the apparent shear
modulus for a bearing may be changed due to bulging and stability
considerations.
Buckling of elastomeric bearings (19,31,56) is dominated by shear
buckling rather than the flexural buckling commonly encountered in columns. Haringx (31) performed the initial work in this area, and showed
that
4P E
1+ - 1
Pcr

PG

(U)

2
P,G

in which PE = Euler buckling load; and PG = GAS in which As = the


effective shear area. Elastomeric bearings are relatively stiff in flexure
and compression but flexible in shear, and so, Eq. 11 reduces to
Per = V P T P G

(12)

Use of equations similar to Eqs. 4 and 6 to calculate the bending stiffness


of the bearing provides a lower bound on the buckling stress of
irGLS

in which h = the total height of the bearing. This relation may be very
conservative in some cases but has the virtue of simplicity.
As will be examined later, fatigue and tensile crack growth in the elastomer are major concerns in the design of elastomeric bearings. Crack
growth theory was developed and is widely accepted for linear elastic
metals, but Rivlin and Thomas extended this theory to rubber, and other
studies (6,35,38,41,42,43) have considered crack growth as a potential
failure mechanism for elastomeric bearings. Calculations of required energy levels have been made and compared with experimental results.
The results indicate that fatigue cracks in bearings under compression
are initiated at tensile stress concentrations at edges of the bond between
rubber and reinforcement. Ozone or mechanical cracking are common
causes of stress concentrations. This illustrates the importance of using
edge cover to protect the sensitive edge of the bond and so improve
fatigue life. Careful shaping (such as addition of radiused fillets) has
been shown to provide further improvement but is seldom used because
of manufacturing difficulties. Crack propagation has gained wide acceptance for elastomers, but the acceptance is not universal (58). This is
one area where existing knowledge is incomplete.
2859

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

EXPERIMENTAL VERIFICATION

Nearly all of the foregoing theories are based on assumptions or hypotheses which are subject to debate. Experimental evidence usually resolves much of this debate. However, in the case of elastomeric bearings, frequent conflicts exist in the experimental evidence. These are
caused in part by inevitable variations in material properties, but are also
increased by errors and improperly designed experiments. Thus, only
the most fundamental questions relating to elastomeric bearings are free
from conflicting evidence.
The earliest tests on elastomeric bearings were primarily concerned
with behavior under compressive loads. The Dupont Company (15) performed an early series of tests on unreinforced pads, and they generated
empirical load-deflection curves for pads of different hardnesses and shape
factors. This work apparently serves as a basis for the existing American
Association of State Highway and Transportation Official (AASHTO)
Specification (57), and the same firm has recently coordinated similar
tests (18) on full size bearings. Numerous other compression tests
(9,12,13,20,47,55) have also been performed. The results of these tests
clearly show that the force-deflection behavior is highly nonlinear as
shown in Fig. 5, but there is wide scatter of the actual numerical results.
Thus, there are several methods (10,15,20) for predicting bearing stiffness with little reason (58) to favor one over the others. There are several
reasons for this scatter, but variations in the experimental methods such
as rate of loading and deflection measurement will have significant impact. Rapid loading and measurement will record some of the elastomer
stiffening associated with dynamic loading. Slow loading and measurement will likely record part of the creep and other long term deformation. However, these features are frequently not recognized or reported.
The results consistently show that bearings with harder elastomer and
higher shape factors are stiffer. Some of these bearings (10,12,18,55) were

FIG. 5.Typical Force Deflection of


Elastomeric Bearings in Compression

FIG. 8.Roll Over at Edge of Bearing

2860

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

also monotonically loaded to failure. The bearings with properly designed steel reinforcement developed high ultimate loads. Maximum average compressive stress, a max , was nearly always above 6 ksi (41 MPa)
and sometimes as high as 18 ksi (120 MPa). The ultimate failure was
nearly always by tensile failure of the steel reinforcement. Fiberglass
reinforced bearings failed at much lower loads since the reinforcement
was substantially weaker. This clearly shows the importance of reinforcement strength to the capacity of the bearing, and it indicates that
steel reinforced bearings have considerable reserve strength under
monotonic loading. However, it should be noted that yield of the reinforcement occurred at loads on the order of 25% of the ultimate load
and so a factor of safety of at least 4.0 is needed against this failure load.
Holes in the reinforcement cause stress concentrations and reduce the
net section and, thus, significantly reduce ultimate load capacity. However, structural designers and bearing manufacturers frequently put holes
in reinforcement with no thought to such consequences.
Bearings have also been tested under shear (12,18,47,50,55) and rotation (52) (as shown in Figs. 3(fr)-3(c)). There is agreement that shear
behavior can be adequately modeled by Eq. 10 if it is recognized that G
increases for dynamic loadings and low temperatures. The magnitude
of the increase varies greatly and can be reliably predicted only by testing the finished bearing. In addition, it should be noted that shear loading causes roll-over of the corners, as shown in Fig. 6 when the bearing
is subjected to large strains. This local effect appears to have only a small
influence on the adequacy of Eq. 10, but it may cause deformation or
other problems if the reinforcement is too thin.
Very limited testing (10,52) has been performed with rotation on bearings. These tests suggest that rotation is a most critical criterion for governing the strength of bearings. Small rotations may cause very large
shear strains in the elastomer, and thus greatly reduce the strength of
a bearing. Further, combined loads on bearings such as shear with
compression and shear with rotation are not well understood from existing research results. Therefore, more experimental research is warranted in these areas.
Other tests (19,20,26,56) have been performed to determine the validity of mathematical theories in modeling the true behavior of bearings.
Gent (26) investigated the stress and strain distribution in bearings loaded
with compression and shear loadings, and compared the results to theory. The compression theory (Eqs. 4 through 9) indicates that the true
compressive stress at the load surface must be distributed as a parabola,
and this distribution has been verified by experiments. Shear deformation theory (Eq. 10) requires that shear stress be constant over the loaded
surface, but it was noted earlier that this does not satisfy equilibrium at
the edge of the bearing. Gent's (26) experiments suggest this constant
shear stress distribution is nearly correct, and the edge discontinuity was
so sharp that it could not be measured in the experiment. Experiments
have been performed to verify the validity of the stability theory (Eqs.
11-13). The results indicate that the theory is valid, but it should be
noted that these experiments were performed only on small bearings
with low shape factors. Therefore, additional testing is warranted with
bearings more typical of those encountered in practice.
2861

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

Tensile rupture and fatigue are widely recognized as critical failure


modes of the elastomer, and thus numerous experiments have studied
these effects. Internal rupture (24) will occur within a bearing at locations with relatively low hydrostatic tensile stress. Hydrostatic tensile
stress may occur on the tension side of a bearing subject to rotation, or
in bearings subjected to tensile uplift. It has been suggested that hydrostatic tensile stresses as low as E may induce this rupture. It should
be noted that internal rupture does not mean the bearing loses all capacity to resist increased loading, but it introduces the potential for a
shortened service life. Thus, tensile load on bearings is inadvisable, and
rotations are usually restricted to help control the development of tensile
stress.
Fatigue of elastomeric bearings may occur even if the bearing is continuously in compression because tensile stresses develop due to the
bulging. One of the earliest studies (8) of fatigue in rubber consisted of
repeated compression or shear loading on bonded blocks of strain crystallizing rubber. Increased fatigue life resulted if complete relaxation was
not permitted after each cycle and if strain cycles did not pass through
zero strain. These results suggested that severe strain cycles can be applied with long fatigue life, if a minimum strain is maintained at all times,
and some authorities (16) have extrapolated this finding to suggest that
the fatigue life of elastomeric bearings is related to elongation at break.
Other more recent fatigue tests (7,10,50) have been performed on actual
elastomeric bearings, and some (10) of these results directly contradict
the previously mentioned extrapolation. The tests to date indicate that
cyclic rotation is the most severe fatigue loading, with cyclic shear normally more critical than repeated compression. Limited tests (10) also
suggest that fatigue failure of the steel laminate may be critical when
the reinforcement has holes or other discontinuities. The fatigue tests
described in existing literature frequently are not well documented and
are poorly coordinated. Thus it is not surprising that there are numerous
contradictions. This is clearly an area which requires additional work.
MODES OF FAILUREDESIGN CRITERIA

Some of the major factors relating to the behavior of elastomeric bearings have been examined here. Many of these factors are not familiar to
structural engineers who must design or select a bearing. However, they
should be considered in the design of a bearing, and therefore, it is valuable to summarize potential modes of failure:
1. Fatigue of the elastomer is widely accepted as a prime concern in
the design of bridge bearings, but there is disagreement over how fatigue is controlled. Most tests to date have attempted to find suitable
fatigue or endurance limits for bearings under various loadings, but the
results of these tests are contradictory and inconclusive.
2. Failure of the reinforcement due to rupture or yield under monotonic load or fatigue under cyclic loading is also a serious problem. This
problem is compounded by holes or discontinuities in the reinforcement.
3. Stability failure may occur for surprisingly short bearings because
of the low shear stiffness of rubber.
2062

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

4. Excessively stiff bearings induce large expansion forces in bridge


piers and the attached substructure, which may cause damage to the
structure. This is a particular problem with low temperature or dynamic
loadings, since elastomers may stiffen greatly under these conditions.
5. Serviceability failure may occur if the bearing deflects too much in
compression due to excess flexibility or creep. Excess deflections may
damage the structure or render it unusable.
6. Bond failure or delamination of the reinforcement has historically
been a major source of problems in bearings, but today it has been largely
eliminated as a problem provided the bearing is vulcanized after bonding and reasonable standards of cleanliness and quality control are
followed.
7. Internal rupture and tension cracking of the elastomer are also
possible modes of failure, but they are frequently controlled by limiting
the magnitude of loadings or deformations which may cause tensile stress
or strain.
8. Elastomers are virtually indestructible under hydrostatic compressive loading and normal environmental conditions, but extremes such
as very low temperatures or high chemical concentrations may cause
deterioration of the bearing.
9. Slippage of the bearing is also a critical problem, particularly with
unreinforced pads.
10. Finally, it must be noted that there are limitations in the present
understanding of elastomeric bearings. For example, the effects of combined loads or unusual geometries are not well understood. Thus, designers must be particularly cautious when dealing with new or unusual
conditions.
DESIGN METHODS

A number of design methods for elastomeric bearings have evolved.


Nearly every major country has a different specification with a different
design rationale. These differences are very significant. However, each
of these specifications can usually be traced back to one of four different
design methods. The four basic methods are quite different and produce
very different designs, but they all attempt to satisfy the design criteria
described here. This section will present a review and comparison of
these methods.
The 1977 AASHTO Specification (57) typifies one of these methods.
This is a simple, empirical design method which was based on some
early industrial research (15) in the United States. It tends to be very
conservative in the design of reinforced bearings, but it allows the designer and bearing manufacturer considerable latitude without requiring
extensive analysis. It tends to be somewhat unconservative for unreinforced pads since they are designed in the same way as reinforced bearings. It limits a c , the average compressive stress, to 800 psi (6 MPa) for
combined dead and live load, and 500 psi (3 MPa) for dead load only.
The average compressive strain, lc, is limited to 0.07. These limitations
are not totally rational, but they perform a valuable function since they
help to assure satisfactory fatigue life of the bearing (58). Stability is assured by limiting the total thickness to a proportion of the plan dimen2863

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

sions. The maximum shear deflection between extreme conditions is 0.5


T, in which T = the total thickness of the elastomer. There is no consideration or limitation of rotation in the specification, but standard practice limits rotation to less than 2AC/L and 2AC/W, in which Ac = the compressive deflection (Fig. 3). Slippage is limited by requiring that the bearing
be restrained against slip if the average compressive stress is ever below
200 psi (1 MPa). The strength of the reinforcement is not considered.
The German procedure (30) is also a simple empirical method, but
beyond this it is the antithesis of the AASHTO approach. There is no
formal specification. Instead, a formalized contractual agreement with
each manufacturer is formed. This agreement permits only a single elastomer (chloroprene) compound and requires rigid quality control during
the manufacturing process. Although the concept of a shape factor is
not used, the typical sizes result in shape factors which are nearly always between 7 and 13. The allowable stress on the bearings is limited
to 1,420, 1,775, and 2,130 psi (10, 12, and 15 MPa) with a 50% increase
permitted it shear deformations are limited with a sliding device. Shear
deformations are limited to less than 0.6T or 0.7T. Stability is assured
by limiting total rubber thickness to a proportion of the plan dimensions.
Unreinforced pads are not permitted. The preceding description permits
selection of a bearing, but it is not the key to German practice. This
consists of extensive certification testing which is required before a manufacturer produces bearings. This certification requires approximately a
year of testing and involves considerable cost. Further, periodic recertification and independent quality control supervision are required. As
a result, only four manufacturers are presently certified to practice.
The British Specification BE 1/76 (16) is probably the most widely used
elastomeric bearing specification in the world. It is based on extensive
research (20,21,24,25) performed in the late 1950s. The method recognizes that shear strains, yc and ys (Fig. 3), may be additive and they are
limited so that
Is = j

s 0.5

(14)

and yc + ys^i]EB

(15)

in which EB = the elongation at break of the elastomer compound (Fig.


7); and i) = a ratio (1/3 or 1/2) which is a function of loading. The elongation at break limitation is based on an extrapolation of an early fatigue
study (8). The shear strain due to compression, yc, is calculated from
lc = 6 S i

, . ,
in which

(16)
P
e=

(17)

y
AeE(l + 2kS2)
'
k = a given dimensionless material constant; and Ae = the effective area
of the bearing. These equations are based on the approximate theoretical
work of Gent and Lindley (25). The reduced effective area (Fig. 8) is an
empirical concept which is inserted to account approximately for combined loadings and the P-A effect. Rotations are limited as is the standard practice with AASHTO, but no effort is made to include shear strain

2864

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

1.

2.

3.

<l.

ELONGATION ( I N / I N )

FIG. 7.Typical Uniaxial Stress Strain


Curve for Rubber

FIG. 8.Effective Area for Bearings


Loaded with Combined Shear and
Compression

due to rotation even though these strains may be large. Stability is assured by limiting the total rubber thickness to a proportion of the plan
dimensions, but this proportion is much less conservative than AASHTO.
The reinforcement is designed by a simple equation which is based on
linear elastic compression theory (58). Unreinforced pads are treated more
conservatively than reinforced bearings, since strains, yc and e, deflection, Ac, and shape factor, S, are increased by 1.8 to account for slip.
The fourth design method is based on the international railway specification (Union Internationale des Chemins de Fer) UIC772R (13). The
specification is based on an extensive research program (10) on full size
bearings, but the research is not well documented. Its primary design
limitations are
(18)

Tc + Ts + Tr S 5 G .

ys = - < 0.7
(19)
'
G
Shear stresses, TC , TS , and j r , are limited rather than shear strains as in
BE 1/76 (16). However, this difference is not overly significant since shear
stress and strain are directly related. The method is different in that it
includes shear stress due to rotation, Tr, which is very significant and
it limits the total to 5G. The 5G limitation is an empirical result based
on fatigue tests (10), and it is a direct contradiction to the BE 1/76 (16)
procedure. This specification also recognizes that the shear modulus increases with dynamic loads, and thus dynamic loadings are penalized:
and

1.5(PP + 1.5Pd)
SAe

(20)
2865

J. Struct. Eng., 1983, 109(12): 2853-2871

_..
c 30

AASIITO
.

BE 1/76

. ' UIC 772 R

STANDARD GEfflAH SIZE

HEAR

1
55

>SIVE STRES

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

. . _

I
1.

Jb

/0

s'

//
/

<XJO@

O/ "6
O
E, - H

'

FIG. 9.Comparison of Design Capacity of Various Codes with No Shear or Rotation for Reinforced Bearings

--AASHT0
.

BE 1/76

s?

UIC 772 R

V
&

^ '/ _ ^ i " E, - 600!


Ts-'_____^
.//
~-"

>

E, - 350!

in
2T(L<)

FIG. 10.Comparison of Design Capacity with Maximum Shear and Rotation for Reinforced Bearings

FIG. 11.Comparison of Design Capacity of Unreinforced Pads

= 2 G + Gp.

(21)

GL2

(22)

in which a = the rotation; P = the compressive load; and subscripts p


2866

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

and d indicate static and dynamic loads. These equations also were approximately derived from theory and thus have some rational basis (58).
The rotation and compression are further limited so that
6AC
tan ap + 1.5 tan ad <

(23)

L-i

and

* _ P_ * 2LGS
_ . . . .

(24)

It should be noted that the average compressive stress, a, is also frequently limited to 1,450 psi (10 MPa) in practice, but this limitation is
not included in the specification. Stability is assured by an equation similar to Eq. 13, and reinforcement is treated in a manner similar to that
in BE/176. Unreinforced bearings are also designed more conservatively
due to the possibility of slip and
a < 2GS

(25)

A detailed analysis of these specifications is presented in another report (58), but this brief comparison clearly indicates that a wide range
of design procedures have been used for elastomeric bearings. Figs. 911 compare the allowable compressive stress allowed by these specifications on reinforced bearings with no shear or rotational deformation
(Fig. 9), reinforced bearings with maximum permissible shear and rotation (Fig. 10), and unreinforced pads (Fig. 11). The comparisons are
made for an elastomer of approximately 55 hardness (G - 110 psi or
0.76 MPa) with elongations at break of 350% and 600% since these are
typical of the material used or specified throughout the world. The bearings are assumed to be square, although a rectangular configuration is
best when rotations are large. Finally, the live loads and dead loads are
taken to be of equal magnitude.
This comparison clearly shows that United States practice as typified
by the 1977 AASHTO Specification (57) is very conservative for reinforced bearings with large shape factors and no shear deformation or
rotation, and UIC772R is very generous in these same areas. United States
practice is much less conservative for small shape factors and bearings
with significant shear and rotation, and it may be regarded as somewhat
unconservative for unreinforced pads. The BE 1/76 Specification is very
dependent on the elongation at break of the elastomer. If it is regarded
as the most rational specification, one can readily see that AASHTO is
not overly conservative with a minimum elongation at break of 350%.
Finally, a comparison of Figs. 9-10 shows the importance of shear deformation and rotation on the load capacity of the bearing. However,
AASHTO totally neglects this effect and only UIC772R fully considers
both. Strength reduction due to stability is not shown on these curves,
but BE 1/76 is generally the least conservative with respect to stability.
In viewing the wide variations in design procedures and design results, it is tempting to draw the conclusion that the design procedures
do not matter. However, this would be erroneous. The AASHTO specification is a simple, conservative design procedure for reinforced bearings, with only minimum standards of acceptance of the material and
2867

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

of manufacturer's quality control. Many U.S. designers make relatively


brash assumptions during the design. For example, they sometimes neglect all rotation or shear movement, or both. This is unrealistic because
bearings are not needed if movements do not occur, but the designer is
protected during this assumption by the concervative code. Other countries may use a much higher allowable stress, but they also require greater
design sophistication, more extensive calculations, and greater quality
control on the material and manufacturing process.
SUMMARY AND CONCLUSIONS

This has been a brief summary of the worldwide state-of-the-art for


elastomeric bearings. A more detailed presentation is given elsewhere
(58). Structural engineers sometimes misunderstand or misinterpret the
behavior and design of elastomeric bearings, and it is hoped that this
paper will help resolve some of these problems. There are deficiencies
in the present design specification for elastomeric bearings, and proposed modifications to correct these differences are also presented elsewhere (58) by the writers.
Finally, it must be noted that there are numerous contradictions in the
worldwide understanding of elastomeric bearings, and research is needed
to resolve them. Some major areas of needed research are:
1. A better resolution of the modes of failure of elastomeric bearings
(particularly fatigue) with consistent guidelines for preventing failure.
2. A better understanding of the effect of rotation and combined loadings on the bearing, and a translation of this understanding into improved design codes.
3. A better understanding of material behavior with emphasis on concerns such as low temperature stiffening.
ACKNOWLEDGMENTS

This work was performed with the support of the National Cooperative Highway Research Program (NCHRP) of the National Research
Council under the guidance of Dr. Robert J. Reilly and NCHRP Project
Panel 10-20. The statements expressed in this paper do not necessarily
reflect the opinions of AASHTO, NCHRP, or the Federal Highway
Administration.
The writers wish to thank NCHRP and the Project Panel 10-20 for their
support and guidance during this study. They also want to express their
gratitude to former graduate students, K. Gottleaber and S. Demitriou,
for their help during the study. The writers are also particularly grateful
to the members of their Advisory Panel, consisting of C. Dolan, J. Kelly,
P. Lindley, L. Mayo, A. Shloss, G. Stranaghan, U. Vasishth, and the
late L. Bell, for their helpful advice and untiring assistance throughout
the study.
APPENDIX.REFIRENCES

1. Aldridge, U. U., Sestak, J. J., and Fears, F. K., "Tests on Five Elastomeric
2868

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.

13.
14.
15.
16.
17.

18.
19.
20.
21.
22.

Bridge Bearing Materials," Highway Research Record, No. 253, 1968, pp. 7283.
Bakirzis, B. A., and Lindley, P. B., "Slipping at Contact Surfaces of Plain
Rubber Pads in Compression," Civil Engineering and Public Works Review, Vol.
65, No. 764, Mar., 1970, pp. 306-307.
Braden, M., and Gent, A. N., "The Attack of Ozone on Stretched Rubber
Vulcanizates. I. The Rate of Cut Growth," Journal of Applied Polymer Science,
Vol. 3, No. 7, 1960, pp. 90-99.
Braden, M., and Gent, A. N., "The Attack of Ozone on Stretched Rubber
Vulcanizates. II. Conditions for Cut Growth," Journal of Applied Polymer Science, Vol. 3, No. 7, 1960, pp. 100-106.
Braden, M., and Gent, A. N., "The Attack of Ozone on Stretched Rubber
Vulcanizates. III. Action of Antiozonants," Journal of Applied Polymer Science,
Vol. 6, No. 22, 1962, pp. 449-455.
Breidenbach, R. F., and Lake, G. J., "Application of Fracture Mechanics to
Rubber Articles, Including Tyres," Philosophical Transactions of the Royal Society
of London, A299, 1981, pp. 189-202.
Breilmaier, A. A., and Hoblitzell, J. R., "Neoprene Bearing Pads Under Repeated Compression and Shear," Highway Research Record, No. 253, 1968, pp.
97-104.
Cadwell, S. M., Merrill, R. A., Sloman, C. M., and Yost, F. L., "Dynamic
Fatigue Life of Rubber," Industrial and Engineering Chemistry, Analytical Edition, Vol. 12, No. 1, Jan., 1940, pp. 19-23.
Clark, E. V., and Moultrop, K., "Load-Deflection Characteristics of Elastomeric Bridge Bearing Pads," Highway Research Record, No. 34, 1963, pp. 90116.
Code 772R, "Code for the Use of Rubber Bearings for Rail Bridges," Union
International des Chemins der Fer, 1973.
Conversy, F., "Appareils d'Appui en Caoutchouc Frette," Annales des Fonts
et Chaussees, IV, Nov.-Dec, 1967.
Crozier, W. F., Stoker, J. R., Martin, V. C , and Nordlin, E. F., "A Laboratory Evaluation of Full Size Elastomeric Bridge Bearing Pads," Transportation Laboratory Research Report CA, DOT, TL-6574-I-74-26, Highway Research Report, June, 1979.
D-60 Report (Interim Reports 1 and 2 and Final Report with Appendices),
Union International des Chemins der Fer, Oct., 1965.
Derham, C. J., "Creep and Stress Relaxation of RubbersThe Effects of Stress
History and Temperature Changes," Journal of Materials Science 8, 1973, pp.
1023-1029.
"Design of Neoprene Bearing Pads," Dupont, Apr., 1959.
"Design Requirements for Elastomeric Bridge Bearings," Technical Memorandum (Bridges) No. BE 1/76, Department of Environment, Highways
Directorate.
"Draft British Standard BS5400: Steel, Concrete and Composite Bridges: Part
9A: Code for Practice for Design of Bearings and Part 9B: Specification for
Materials, Manufacture and Installation of Bearings," Document 81/10184,
British Standards Institution, Feb. 25, 1981.
"Dupont Elastomers for Neoprene Bearings," Bulletin 3A, Dupont, Sept.,
1981.
Gent, A. N., "Elastic Stability of Rubber Compression Springs," Journal of
Mechanical Engineering Science, Vol. 6, No. 4, 1964, pp. 318-326.
Gent, A. N., "Load-Deflection Relations and Surface Strain Distributions for
Flat Rubber Pads," Rubber Chemistry and Technology, Vol. 31, 1958, pp. 395-=
414.
Gent, A. N., "On the Relation Between Indentation Hardness and Young's
Modulus," Transactions of the Institution of the Rubber Industry, Vol. 34, No. 2,
1958, pp. 46-57.
Gent, A. N., "Relaxation Processes in Vulcanized Rubber. I. Relation Among
Stress Relaxation, Creep, Recovery and Hysteresis," Journal of Applied Poly2869

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

mer Science, Vol. 6, No. 22, 1962, pp. 433-441.


23. Gent, A. N., "Relaxation Processes in Vulcanized Rubber. II. Secondary Relaxation Due to Network Breakdown," Journal of Applied Polymer Science, Vol.
6, No. 22, 1962, pp. 442-448.
24. Gent, A. N., and Lindley, P. B., "Internal Rupture of Bonded Rubber Cylinders in Tension," Royal Society of London Proceedings, Series A, Vol. 249, 1959,
pp. 195-205.
25. Gent, A. N., and Lindley, P. B., "The Compression of Bonded Rubber Blocks,"
Institution of Mechanical Engineers Proceedings, Vol. 173, pp. 111-222.
26. Gent, A. N., Henry, R. L., and Roxbury, M. L., "Interfadal Stresses for Bonded
Rubber Blocks in Compression and Shear," Transactions of the ASME, Journal
of Applied Mechanics, Vol. 96, No. 4, Dec, 1974, pp. 855-859.
27. Gent, A. N., and Meinecke, E. A., "Compression, Bending and Shear of
Bonded Rubber Blocks," Polymer Engineering and Science, Vol. 10, No. 1, Jan.,
1970, pp. 48-53.
28. Gregory, M. J., "Effect of Compounding on the Properties of Natural Rubber
Engineering Vulcanizates," Natural Rubber Engineering Data Sheet EDS3, Malaysian Rubber Producers Research Association, England, 1979.
29. Gregory, M. J., Metherell, C , and Smith, J. F., "The Effect of Carbon Black
Fillers on Viscoelastic Properties of Vulcanizates," Plastics and Rubber: Materials and Applications, Feb., 1978, pp. 37-40.
30. Gumba-Gummi in Bauwesen GMBH, (Approval by Institute of Civil Engineering for Elastomeric Bearings Produced by GUMBA), West Germany, Oct.,
1973.
31. Haringx, J. A., "On Highly Compressive Helical Springs and Vibration-free
Mountings, Parts I, II, III," Philips Research Reports, 1948-1949.
32. Holownia, B. P., "Compression of Bonded Rubber Blocks," Journal of Strain
Analysis, Vol. 6, No. 2, 1970, pp. 121-123.
33. Holownia, B. P., "Effect of Poisson's Ratio on Bonded Rubber Blocks," Journal of Strain Analysis, No. 7, 1972, p. 236.
34. Holownia, B. P., "Effect of Different Types of Carbon Black on Elastic Constants of Elastomers," Plastics and Rubber: Materials and Applications, Aug.,
1980, pp. 129-132.
35. Lake, G. J., and Lindley, P. B., "Ozone Attack and Fatigue Life of Rubber,"
Use of Rubber in Engineering, Maclaren & Sons Ltd., London, England, 1967,
pp. 56-71.
36. Lake, G. J., and Lindley, P. B., "Ozone Cracking, Flex Cracking and Fatigue
of Rubber," Rubber Journal, Oct.-Nov., 1964, pp. 1-16.
37. Lindley, P. B., "Compression Moduli for Blocks of Soft Elastic Material Bonded
to Rigid End Plates," Journal of Strain Analysis, Vol. 14, No. 1, 1979, pp. 1116.
38. Lindley, P. B., "Ozone Attack at a Rubber-Metal Bond," Journal of the IRI,
Vol. 5, No. 6, Dec, 1971, pp. 243-248.
39. Lindley, P. B., "Plane Strain Rotation Moduli for Soft Elastic Blocks," Journal
of Strain Analysis, Vol. 14, No. 1, 1979, pp. 17-21.
40. Lindley, P. B., "Plane-Stress Analysis of Rubber at High Strains Using Finite
Elements," Journal of Strain Analysis, Vol. 6, No. 1, 1971, pp. 45-52.
41. Lindley, P. B., "Relation Between Hysteresis and the Crack Growth Resistance of Natural Rubber," International Journal of Fracture, Vol. 9, No. 4, Dec,
1973, pp. 449-462.
42. Lindley, P. B., and Stevenson, A., "Fatigue Resistance of Natural Rubber in
Compression," Malaysian Rubber Producers Research Association, England, Mar.,
1981.
43. Lindley, P. B., and Teo, S. C , "Energy for Crack Growth at the Bonds of
Rubber Springs," Plastics and Rubber: Materials and Applications, Feb., 1979,
pp. 29-37.

44. Lindley, P. B., and Teo, S. C , "High Temperature Aging of Rubber Blocks,"
Plastics and Rubber: Materials and Applications, May, 1977, pp. 82-88.
45. Meier, J. R., Rudd, G. E., and Weir, D. F., "Creep Resistant Elastomer For2870

J. Struct. Eng., 1983, 109(12): 2853-2871

Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 02/03/16. Copyright ASCE. For personal use only; all rights reserved.

46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.

mulations," Rubber Chemistry and Technology, Vol. 52, No. 1, Mar.-Apr., 1979,
pp. 50-71.
Meinecke, E. A., "Comparing the Time and Rate Dependent Mechanical
Properties of Elastomers," Rubber Chemistry and Technology, Vol. 53, No. 5,
Nov.-Dec, 1980, pp. 1145-1159.
Minor, J. C , and Egen, R. A., "Elastomeric Bearing Research," NCHRP Report 109, National Research Council, Washington, D.C., 1970.
Murray, R. M., and Detenber, J. D., "First and Second Order Transitions in
Neopiene," Rubber Chemistry and Technology, Apr.-June, 1961.
Nachtrab, W. B., and Davidson, R. L., "Behavior of Elastomeric Bearing Pads
Under Simultaneous Compression and Shear Loads," Highway Research Record, No. 76, 1965, pp. 83-101.
Ozell, A. M., and Diniz, J. F., "Report on Tests of Neoprene Pads Under
Repeated Shear Loads," Highway Research Record, Vol. 242, 1960, pp. 20-27.
Porter, L. S., and Meinecke, E. A., "Influence of Compression Upon the
Shear Properties of Bonded Rubber Blocks," Rubber Chemistry and Technology,
Vol. 53, No. 5, Nov.-Dec, 1980, pp. 1133-1144.
Price, A. R., "Abnormal and Eccentric Forces on Elastomeric Bridge Bearings," TRRL Laboratory Report 708, 1976.
Rejcha, C , "Design of Elastomer Bearings," PCI Journal, Vol. 9, No. 5, Oct.,
1964, pp. 62-78.
Rivlin, R. S., and Saunders, D. W., "Cylindrical Shear Mountings," The British Rubber Producers Research Association, Publication No. 115.
Sanpaolesi, L., and Angotti, F., "Appareils D'Appui en Caoutchouc pour les
Constructions," (Rubber Bearings for Construction), Construction Metallique,
No. 1, 1972, pp. 21-42 (in French).
Schapery, R. A., and Skala, D. P., "Elastic Stability of Laminated Elastomeric
Columns," International Journal of Solids and Structures, Vol. 12, No. 6, 1976,
pp. 401-417.
Standard Specifications for Highway Bridges, American Association of State
Highway and Transportation Official, 12th Edition, Washington, D.C., 1977.
Stanton, J. F., and Roeder, C. W., "Elastomeric BearingsDesign, Construction and Materials," NCHRP Report 248, Sept., 1982.
Topaloff, B., Gummilager fur Briicken-Berechnung und Anwendung," (Rubber Bearings for BridgesCalculations and Uses), Der Bauingenieur, Vol. 39,
No. 2, Feb., 1964, pp. 50-64 (in German).
Treloar, L. R. G., The Physics of Rubber Elasticity, Clarendon Press, Oxford,
England, 1975, p. 310.

2871

J. Struct. Eng., 1983, 109(12): 2853-2871

You might also like