Anda di halaman 1dari 12

Shear and Friction Response of Nonseismic Laminated

Elastomeric Bridge Bearings Subject to Seismic Demands


Joshua S. Steelman, S.M.ASCE1; Larry A. Fahnestock, P.E., M.ASCE2; Evgueni T. Filipov, S.M.ASCE3;
James M. LaFave, P.E.4; Jerome F. Hajjar, P.E., F.ASCE5; and Douglas A. Foutch, P.E., M.ASCE6
Downloaded from ascelibrary.org by CRRI - Central Road Research Institute on 01/09/18. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Laminated elastomeric bridge bearings are commonly used in areas with low-to-moderate seismicity, although the applications are
typically intended for service-level considerations such as thermal movements of the bridge superstructure. These components provide a po-
tential source of displacement capacity frequently neglected in seismic design. An experimental program was carried out to evaluate the behav-
ioral characteristics and performance of steel-reinforced, laminated elastomeric bearings, which had not been designed for seismic demands, as
the primary quasi-isolation components for seismic events by permitting slip at the interface of the bearing and substructure. The rubber at the top
of the bearing is vulcanized to a steel plate, which is bolted to the test frame to simulate a connection to the superstructure. At the base of the
bearing, the elastomer directly contacts concrete representing the substructure, with no restraint of horizontal motion other than friction. The
elastomeric bearings investigated during the experimental program displayed an approximately linear elastic response before sliding, with an
initial friction coefficient in the range of 0.25–0.5 (at a shear strain between 125 and 250%) depending on combinations of the contact surface
roughness, applied load, and bearing velocity. The friction coefficient decreased as a nonlinear function of the imposed vertical load. The maximum
elastomer shear strain prior to sliding exhibited nonlinear increases with vertical load, resulting from the influence of the variable friction coefficient.
Linear shear moduli were primarily influenced by the maximum shear strain imposed on the bearing, and showed shear stiffness reductions of
approximately 40–50% following multiple, large displacement slip cycles, compared with 15–25% after reaching 50% shear strain. Multiple cycles
of large displacement demands resulted in noticeable degradation in the friction coefficient over the duration of the tests. However, the bearings
possessed a high degree of resiliency, considering that the specimens retained load-carrying capacity through total cumulative slip travel demands
in the range of 3.5–4.5 m (140–180 in.). DOI: 10.1061/(ASCE)BE.1943-5592.0000406. © 2013 American Society of Civil Engineers.
CE Database subject headings: Highway bridges; Earthquake resistant structures; Friction; Cyclic tests; Seismic effects; Laminated
materials; Shear resistance.
Author keywords: Elastomeric bearings; Quasi-isolation; Full-scale tests; Friction coefficient.

Introduction of the earliest patents dating back to the turn of the twentieth century
(Buckle and Mayes 1990). Extensive subsequent research efforts
Seismic isolation has been implemented as an effective means of have further developed seismic isolation design methods and related
earthquake damage mitigation in structural engineering, with some bearing technologies (e.g., Calvi and Pavese 1997; Kikuchi and Aiken
1997; Mayes and Naeim 2001; Buckle et al. 2002; Taylor and Igusa
1
Graduate Research Assistant, Dept. of Civil and Environmental Engi- 2004). Structural isolation systems have typically been used in high
neering, Univ. of Illinois at Urbana–Champaign, 205 N. Mathews Ave., seismic regions (such as the west coast of the United States, as well as
Urbana, IL 61801. E-mail: jsteelm2@illinois.edu
2 Japan and New Zealand) where additional design and construction costs
Assistant Professor, Dept. of Civil and Environmental Engineering,
Univ. of Illinois at Urbana–Champaign, 205 N. Mathews Ave., Urbana, IL are justifiable when balanced against the seismic hazard. However, for
61801 (corresponding author). E-mail: fhnstck@illinois.edu bridges outside of high seismic areas some commonly used structural
3 components that permit relative thermal superstructure displacements
Graduate Research Assistant, Dept. of Civil and Environmental Engi-
neering, Univ. of Illinois at Urbana–Champaign, 205 N. Mathews Ave., with respect to the substructure can provide improved seismic perfor-
Urbana, IL 61801. E-mail: filipov1@illinois.edu mance. For example steel-reinforced, laminated elastomeric bridge
4
Associate Professor, Dept. of Civil and Environmental Engineering, bearings may inherently possess properties suitable for an isolation
Univ. of Illinois at Urbana–Champaign, 205 N. Mathews Ave., Urbana, IL
system, even though elastomeric bearing designs in such regions are
61801. E-mail: jlafave@illinois.edu
5
Professor and Chair, Dept. of Civil and Environmental Engineering, nominally restricted to 50–70% shear strain to ensure satisfactory long-
Northeastern Univ., 360 Huntington Ave., Boston, MA 02115. E-mail: jf. term performance (AASHTO 2008a; Stanton and Roeder 1982). Re-
hajjar@neu.edu cently, Konstantinidis et al. (2008) investigated the behavior of such
6
Professor Emeritus, Dept. of Civil and Environmental Engineering, bearings, recognizing that they can sustain shear strains much larger than
Univ. of Illinois at Urbana–Champaign, 205 N. Mathews Ave., Urbana, IL typically observed under nonseismic service-level conditions.
61801. E-mail: dfoutch@illinois.edu Consequently, a quasi-isolated design that relies on thermal ex-
Note. This manuscript was submitted on August 26, 2011; approved on
pansion devices to also provide seismic energy dissipation could
April 24, 2012; published online on April 26, 2012. Discussion period open
until December 1, 2013; separate discussions must be submitted for in-
be an appealing pragmatic design approach for bridges in moderate
dividual papers. This paper is part of the Journal of Bridge Engineering, seismic regions, providing the benefit of reduced force demands,
Vol. 18, No. 7, July 1, 2013. ©ASCE, ISSN 1084-0702/2013/7-612–623/ as long as the structural system can be designed to account for
$25.00. increased displacements associated with an isolated response

612 / JOURNAL OF BRIDGE ENGINEERING © ASCE / JULY 2013

J. Bridge Eng., 2013, 18(7): 612-623


potentially having very low stiffness after sliding is initiated. The characteristics for bearings positively connected to the superstruc-
key characteristics of a quasi-isolated bridge are that the design ture and permitted to slide on the concrete substructure, a bearing
approach is significantly simplified and the fabrication costs for the configuration that has only been studied to a limited degree in prior
bearings are less than conventional seismic isolation bearings. Here, experimental research. Primary focus is placed on multiple cycles to
the term quasi-isolated has been chosen because the system response large displacements in order to evaluate the potential for bearing
resembles that of a conventionally isolated system (AASHTO 2010) degradation and associated changes in response properties. The tests
in that the superstructure can move relative to the substructure with are part of a broader experimental program that is also investigat-
minimal restraint. Although the configuration of a quasi-isolated ing elastomeric bearings with Teflon sliding surfaces (IDOT Type II
system is simpler than that of a conventionally isolated system, this bearings), steel fixed bearings with anchor bolts and pintles, and
benefit must be considered along with the limitations. Compared side (transverse) retainers that are typically used in conjunction with
with a conventionally isolated system, there is larger uncertainty in the the elastomeric bearings (LaFave et al. 2013). The experimental
Downloaded from ascelibrary.org by CRRI - Central Road Research Institute on 01/09/18. Copyright ASCE. For personal use only; all rights reserved.

seismic response of a quasi-isolated system and greater damage is results presented in this paper provide new information to un-
expected, primarily constrained to the bearings. However, compared derstand how steel-reinforced, laminated elastomeric bearings that
with a conventional nonisolated system, a quasi-isolated system is have not been designed or fabricated to tolerate seismic demands
expected to sustain less damage at the substructures. For a quasi- will behave under simulated field conditions when subjected to large
isolated bridge, support surfaces at the substructures may need to displacement demands, which include both elastomer shear and
be enlarged to prevent unseating of bearings, and the bridge super- sliding components.
structure could be permanently offset following sliding of the bear-
ings. Calibration of the standards and guidelines recommended for use
when designing a quasi-isolated bridge are in development as part of Testing Program Background
the project in which the experiments reported here were performed.
The Illinois Center for Transportation (ICT), working in con-
Test Setup
junction with the Illinois Department of Transportation (IDOT), is
currently investigating the behavior of quasi-isolated bridge sys- The tests described in this paper were performed in the Newmark
tems, to calibrate and refine the IDOT earthquake-resisting system Structural Engineering Laboratory at the University of Illinois using
(ERS) design and construction methodology. The IDOT ERS phi- the experimental apparatus shown in Fig. 1, which was specifically
losophy relies in part on prescribed sequential fusing (i.e., slip, designed to realistically reflect field conditions for the full-scale
yielding, or rupture) of specific components to limit forces that can bridge bearing test specimens. Vertical loading was imposed on a
be transferred down through the bridge system (Tobias et al. 2008). bearing using a pair of 445 kN (100 kip) capacity actuators reacting
The specific feature of the IDOT ERS addressed in this paper is the against a steel frame secured to the laboratory strong floor with
use of Type I elastomeric bearings, which allow thermal expansion pretensioned anchors. The use of a pair of actuators allowed the
under typical service conditions, to accommodate much larger vertical load to be maintained at a specified target of simulated
earthquake-induced displacements and provide a significant elon- gravity load (and the loading beam to remain level) regardless of the
gation of the structural period (possibly in conjunction with fixed lateral displacement of the bearing. Concrete pads were cast to
bearing fusing). Bearings identified as Type I by IDOT are fabricated simulate typical bridge substructures, including a brushed finish as
with an external top steel plate vulcanized to a steel-reinforced specified by IDOT (2007), to ensure that the frictional response at the
elastomer block, and are differentiated from Type II and III elas- elastomer-to-concrete substructure interface appropriately reflected
tomeric bearings in that they lack a Teflon sliding surface and an expected field conditions. The concrete pads were secured to the
internal shear restrictor pin within the elastomeric block, re- strong floor using 10 pretensioned anchors around the perimeter of
spectively (IDOT 2009). Slip for Type I bearings is anticipated at the the pad.
interface of the bottom surface of the elastomer and the top surface of The primary objective of the testing program was to characterize
the concrete substructure. the structural performance and behavior of elastomeric bearings
The behavior of steel-reinforced, laminated elastomeric bearings subjected to large lateral deformations during seismic events. Based
has been studied extensively for shear strains up to about 75%, on preliminary systems analyses of full bridges, in conjunction with
which are consistent with typical service conditions (Stanton and a review of the literature documenting previous tests of similar
Roeder 1982; Roeder et al. 1987). Schrage (1981) investigated slip elastomeric bearings (Stanton and Roeder 1982; Roeder et al. 1987;
thresholds of neoprene bearings on concrete surfaces. Mori et al. Kulak and Hughes 1992; Mori et al. 1999; Konstantinidis et al.
(1999) performed experiments on the shear response of laminated 2008), a target of 400% shear strain was selected for the testing
rubber bearings, which were not bonded to top or bottom steel plates program. Actual displacements can include both elastomeric shear
and were thus free to roll, to approximately 200% shear strain, with deformations and frictional sliding because the bearings are posi-
and without stops to restrict slip. Konstantinidis et al. (2008) tested tively connected at their top surface but are not anchored at their base
similar bearing configurations as Mori et al. (1999), without top or against horizontal motion at the elastomer-to-concrete substructure
bottom steel plates, and concluded that the ultimate displacement interface. Thus, although the elastomer portion of the bearings tested
capacity was approximately 150–225% shear strain, limited by in the present research is similar to those tested by Konstantinidis
rollover of the bearing to the point where the originally vertical face et al. (2008), the overall behavior is quite different. The bearings
of the bearing becomes horizontal against the supports. It was noted tested by Konstantinidis et al. (2008) were not bonded to top or
that any further imposed displacement would cause slipping of the bottom plates, thus their large-displacement behavior was a com-
bearing, and that the bearing surface may be damaged by repeated bination of shear and rollover (without sliding).
slipping; however, the experimental results did not exhibit cyclic To homogenize descriptions of the testing protocols across
degradation. experiments with variable slip thresholds, deformations are referred
This paper describes laboratory testing of full-scale steel- to herein using the measure of equivalent shear strain (ESS), where
reinforced, laminated elastomeric bridge bearings to study their ESS is the average bearing shear strain that would have been de-
shear and friction response when subjected to simulated seismic veloped at a given displacement if no slip had occurred. The
demands. The objective of the tests is to determine the behavioral largest bearings in the testing program comprised five layers of

JOURNAL OF BRIDGE ENGINEERING © ASCE / JULY 2013 / 613

J. Bridge Eng., 2013, 18(7): 612-623


15.9 mm (5/8 in.) elastomer, for a total height of rubber (hrt ) of elastomer used in the bearing specimens. The hardness measure-
79.4 mm (3.125 in.), and an associated 400% ESS of 317.5 mm ments shown in Table 1 were averages taken using a durometer at
(12.5 in.). A 762 mm (30 in.) stroke, 980 kN (220 kip) capacity quarter points along each side, at the second and fourth layers of the
actuator was used to provide the required 635 mm (25 in.) total elastomer, for each bearing.
horizontal displacement capacity for fully reversed cyclic testing. An example of the initial state of a bearing prior to application of
lateral load is shown in Fig. 1. The top steel plates were bonded to the
elastomer during the vulcanization process, and positive anchorage
Test Specimens
was provided to secure the bearing to the test frame loading beam,
The experimental results presented in this paper are for the speci- which represents a bridge girder, with four 25 mm (1 in.) diameter
mens listed in Table 1. IDOT Size 7-c and 13-c bearings are rep- threaded steel studs in tapped holes in the bearing top plate. This
resentative of those appropriate for use at abutment and pier attachment method is consistent with bridge construction in Illinois,
Downloaded from ascelibrary.org by CRRI - Central Road Research Institute on 01/09/18. Copyright ASCE. For personal use only; all rights reserved.

locations, respectively. All elastomeric bearings were furnished by except that in practice the four threaded steel studs are typically
Tobi Engineering, Inc., located in Glenview, Illinois. The dimen- 19 mm (3/4 in.) diameter (IDOT 2009). The stud size was increased
sions of the test specimens conformed to standards used by IDOT, as to ensure that the studs and surrounding material would remain
specified in their bridge manual (IDOT 2009); they are provided in elastic and that slip would occur only between the elastomer and
Table 1 (in units of mm) for the elastomer layers, steel reinforcing concrete, not between the top plate and the loading beam. Horizontal
shims, and the top steel plate. According to documentation supplied lines were drawn on the sides of the elastomer to indicate the
by the manufacturer, the elastomer used for the bearings was locations of the internal shims, and vertical lines were added to form
composed of polyisoprene (natural rubber) meeting the AASHTO a grid. The elastomer bore directly on the roughened concrete
thermal requirements of Grade 3. Specifications for bearings sup- surface, without any positive horizontal restraint at the simulated
plied to IDOT, require a Shore A hardness of 55 6 5, which cor- substructure interface (see Fig. 2).
responds to a range of estimated shear moduli from 655 to 1,380 kPa
(95 to 200 psi), according to AASHTO (2008a). All of the bearing
Test Procedure
specimens satisfied this hardness requirement. The manufacturer
reported a hardness of 54.3, and a shear modulus, according to All tests simulated longitudinal motion of a bridge superstructure.
ASTM D4014, Annex A (ASTM 2007), of 855 kPa (124 psi) for the Two 127 cm (50 in.) cable-extension position transducers were

Fig. 1. Testing setup elevation and typical specimen

Table 1. Test Specimen Characteristics


Average
Elastomer layer Top plate
IDOT Shape hardness Shim,
Specimen size W (mm) L (mm) t (mm) Quantity hrt (mm) factor (Shore A) t (mm) W (mm) L (mm) t (mm)
1 7-c 305 178 9.53 5 47.6 5.895 50.5 2.38 356 203 38.1
2 7-c 305 178 9.53 5 47.6 5.895 54.7 2.38 356 203 38.1
3 7-c 305 178 9.53 5 47.6 5.895 53.4 2.38 356 203 38.1
4 13-c 508 330 15.9 5 79.4 6.303 54.7 4.76 559 356 63.5

614 / JOURNAL OF BRIDGE ENGINEERING © ASCE / JULY 2013

J. Bridge Eng., 2013, 18(7): 612-623


attached to welded studs at the center of the bearing top plate to make multiple imposed cycles of strain) that may develop in the elastomer
the primary measurements used for control of the tests. Additionally, material response. Material scragging effects were seen to be
four 63.5 cm (25 in.) cable-extension position transducers were at- minimal after the first excursion to a new maximum shear strain
tached to needles inserted in the sides of the elastomer, inset by during the preslip cycles of the tests.
approximately 19 mm (3/4 in.) from the ends of each bearing. Fig. 1 The rate field in Table 2 indicates whether a test was quasi-static
illustrates the attachment points for these transducers on one side of (QS) or performed at an increased strain rate (ISR) to investigate
a bearing. The parameters for each test are summarized in the testing sensitivity in the response characteristics to velocity demands. QS
matrix shown in Table 2. tests were performed by dividing the displacement record into
In Table 2, the protocol field indicates whether a test was substeps that were composed of iterations on the horizontal actuator
monotonic [including direction, (1) or (2), see Fig. 1] or cyclic, and position to converge to a target horizontal position of the bearing top
the test travel field indicates the total travel of the top plate for that plate, followed by iterations of the vertical actuator positions to
Downloaded from ascelibrary.org by CRRI - Central Road Research Institute on 01/09/18. Copyright ASCE. For personal use only; all rights reserved.

particular test, in units of ESS (where 100% 5 1). The displacement maintain the target simulated gravity load. Because the net vertical
record used for fully reversed cyclic testing, as shown in Fig. 3, was force on the bearing and its horizontal position are related in
determined after a review of documentation available for other a coupled fashion to both the horizontal and vertical actuator
prequalification and characterization tests of seismic isolation positions, particularly at large displacements, successive loops were
bearings [Shenton 1996; Highway Innovative Technology Evalu- performed within a substep to adjust the horizontal and vertical
ation Center (HITEC) 1996; AASHTO 2010]. The initial stages actuator positions until both of their target criteria had converged to
included seven cycles each to 25 and 50% shear strain, followed by within acceptable tolerances. The vertical force on the bearing varied
three cycles each at 100, 200, 300, and 400% ESS. The additional by as much as 6 8.9 kN (2 kip), from the target during vertical
cycles at the lower strain levels were imposed to investigate any control iterations of the QS tests.
effect of scragging (i.e., degradation in stiffness resulting from For the QS tests, the velocity of the top plate was determined
by the time required to sample signals and perform calculations
between iterations, and was generally low relative to the hydraulic
capacity of the testing equipment. For example, the average velocity
for Test 3-1 was approximately 0.068 mm/s (0.0027 in./s), or about
0.14% shear strain per second. For the ISR tests, the command
signals from a previous (slower) test were used to preestablish the
signals corresponding to converged states. Those signals were then
mapped to a sinusoidal waveform on a compressed time line so that
the bearing travel would meet a target velocity. Target velocities for
the ISR tests were selected based on the available hydraulic capacity
of the testing setup, and ranged from 1.5 to 6.3 cm/s (0.6 to 2.5 in./s),
or 32–80% shear strain per second. Although the velocities con-
sidered are smaller than expected for seismic demands, the ISR tests
do provide some insight into potential rate effects. Prior research
[Malaysian Rubber Producers’ Research Association (MRPRA)
1980] shows that a strain rate dependency does exist for natural
rubber with respect to the shear modulus but that the shear modulus
Fig. 2. Specimen 1 (Test 1-1) at incipient sliding
is more significantly influenced by peak shear strain than by strain
rate. More recent research by Konstantinidis et al. (2008) agreed

Table 2. Testing Matrix


Cumulative prior Average vertical
Test number Specimen Protocol Test travel (ESS) tests (ESS) Rate stress (MPa)
1-1 1 Monotonic (2X) 4 0 QS 3.45
1-2 1 Monotonic (2X) 4 4 QS 3.45
1-3 1 Monotonic (2X) 4 8 QS 3.45
1-4 1 Monotonic (1X) 4 12 QS 3.45
1-5 1 Monotonic (1X) 4 16 QS 2.59
1-6 1 Monotonic (1X) 4 20 QS 1.38
1-7 1 Monotonic (1X) 4 24 QS 3.45
1-8 1 Monotonic (1X) 4 28 ISR 3.45
1-9 1 Monotonic (1X) 4 36 QS 3.45
1-10 1 Monotonic (1X) 5 56 QS 3.45
2-1 2 Cyclic 141 0 QS 1.38
2-2 2 Cyclic 141 141 QS 5.52
3-1 3 Cyclic 141 0 QS 3.45
3-2 3 Cyclic 141 141 ISR 3.45
4-1 4 Cyclic 141 0 QS 2.65
4-2 4 Monotonic (1X) 4 141 QS 2.65
4-3 4 Monotonic (1X) 4 145 ISR 2.65
Note: ESS 5 equivalent shear strain; QS 5 quasi-static; and ISR 5 increased strain rate.

JOURNAL OF BRIDGE ENGINEERING © ASCE / JULY 2013 / 615

J. Bridge Eng., 2013, 18(7): 612-623


Downloaded from ascelibrary.org by CRRI - Central Road Research Institute on 01/09/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 3. Cyclic quasi-static testing protocol

with this finding, showing a relatively minor influence of strain rate,


which was overshadowed by the influence of peak shear strain.
Ideally, strain rates for a testing program such as reported in this Fig. 4. Specimen 1 force versus displacement
paper would reach those expected for large seismic events; however,
peak strain demand and capturing the behavior of the bearings when
subjected to large cycles with slip were judged to be even more 330 3 508 mm ð13 3 20 in:Þ. The rate of degradation of friction
significant for characterizing the bearings’ behavior during large resistance was more pronounced (seen as a decrease of sliding force
earthquakes. with increasing cycles) for the larger bearing size, which was per-
Standard IDOT designs are required to maintain an average formed on a new concrete pad (and therefore unaffected by any in-
compression stress between 1.38 and 5.52 MPa (200 and 800 psi) on fluence of prior tests). Increasing vertical load correlates to both an
the elastomer plan area as a result of gravity load. Accordingly, the increased slip load and development of a stiffening branch prior to
compression levels selected for the tests, shown in the average slip, as is evident from a comparison of Figs. 5(a and c).
vertical stress field, fall within this range, and the first phase of each The bearings used for this study are specified primarily on the
test applied a simulated gravity load to the bearing specimen. Each basis of Shore A durometer hardness, with a requirement of 55 6 5.
bearing was first subjected to a vertical loading/unloading/reloading Bearing producers will sometimes add reinforcing fillers such as
cycle to obtain data for the compression stiffness of the specimen carbon black to achieve the required hardness, and will also modify
and also to preload the bearing with the target simulated gravity load the composition of the rubber compound according to pro-
before applying lateral displacements. prietary techniques to optimize other performance objectives re-
quired by departments of transportation. Consequently, the damping
of the rubber compound is not intentionally low or high, and is
Test Results generally ignored when evaluating the structural suitability of the
bearings. When effective damping was calculated for the bearing
Observed Force-Displacement Response at High tests reported in this paper according to AASHTO (2010), the values
Shear Strains obtained for preslip cycles ranged from about 5 to 10%, with a mean
of 7% of critical, for Tests 2-1 and 4-1 with relatively light com-
All bearings exhibited stable hysteretic response at high shear pression loads and slip initiating between 100 and 200% ESS. Under
strains, with modest degradation of slip force levels upon increasing higher compression in Tests 2-2, 3-1, and 3-2, slip was delayed until
numbers of slip cycles. Constitutive force-displacement response after the 200% ESS cycles, and the effective damping ranged from
plots are provided in Figs. 4 and 5(a–d) for selected tests to illustrate about 9 to 18%, with a mean of 13% of critical.
key aspects of their behavior. In Figs. 4 and 5(a–d), points where slip Although delamination of internal shims and instability have
is regarded to initiate are indicated with circular markers. Fig. 4 been observed in previous research on elastomeric bearings (Roeder
presents data for all of the monotonic tests performed on Specimen 1. et al. 1987; Buckle et al. 2002), neither of these failure mechanisms
Successive tests in each direction resulted in degraded stiffness was observed for any of the specimens. Taken as a whole, the results
and slip resistance; however, the incremental stiffness degradation suggest that unrestrained, steel-reinforced, laminated elastomeric
effects diminished with increasing number of cycles. The rapid sta- bearings that have not been specifically designed to resist seismic
bilization of stiffness degradation, with little degradation occurring demands may be permitted to slide on substructures and can be
after the second test in each direction, suggests that the more sig- expected to perform satisfactorily without compromising the
nificant decrease in stiffness from the first test to the second tests may structural integrity of the bearing during seismic events with large
be a consequence of chamfering at the leading edges of the base of the magnitudes, long durations, and multiple cyclic reversals.
elastomer that occurs when the elastomer is ground against the
concrete surface, rather than by unrecovered scragging of the elas-
Elastomeric Bearing Stiffness Characteristics
tomeric material. Tests at increased strain rates resulted in a significant
increase in breakaway coefficient, as well as a minor increase in Shear stiffness of bearings in nonseismic applications is usually
stiffness, which was typically sufficient to overcome the degradation a secondary concern—where the AASHTO specifications are used,
effects of previous tests, as in the case of Test 1-8 in Fig. 4. Method A is often used for bearing designs. Consequently, bearings
Figs. 5(a–c) were obtained from a bearing with an elasto- may be specified for manufacture using only durometer hardness
mer footprint measuring 178 3 305 mm ð7 3 12 in:Þ, whereas limits rather than the shear modulus, as would be required in
the footprint for the bearing represented in Fig. 5(d) is Method B, and the shear modulus used for design capacity checks is

616 / JOURNAL OF BRIDGE ENGINEERING © ASCE / JULY 2013

J. Bridge Eng., 2013, 18(7): 612-623


Downloaded from ascelibrary.org by CRRI - Central Road Research Institute on 01/09/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 5. Force versus displacement results for the cyclic tests: (a) Specimen 2 (Test 2-1); (b) Specimen 3 (Test 3-1); (c) Specimen 2 (Test 2-2);
(d) Specimen 4 (Test 4-1)

assumed based on historically observed correlations with hardness. when the range of the regression was considered together with
Quantification of shear stiffness varies between alternative recog- the loading rate, the increased stiffness was about 12–14% higher
nized standards, where both ASTM (2007) D4014 and AASHTO than would be obtained from AASHTO (2008b) or ASTM (2007)
(2008b) M251 constrain the range of shear strain imposed on test procedures.
specimens to the range of typical design limits but with maximum A general relationship has been recognized between compression
strain demands of 50 and 65%, respectively. and shear stiffness in the form of
When bearings are subjected to seismic demands, shear strains
can extend significantly beyond the nominal service load limits. The s
ɛ¼ (1)
tests performed for this experimental program indicate that elasto- aGS2
mer shear strains could be expected to reach approximately
125–250% before slip, depending on the vertical load. A pro- where ɛ 5 vertical strain, s 5 average vertical stress, G 5 shear
nounced, high-strain stiffening branch associated with crystalliza- modulus, S 5 shape factor, and a 5 constant (primarily influenced
tion was not observed because slip precluded such a phenomenon for by geometry and boundary conditions). Roeder et al. (1987) pre-
each bearing test. Therefore, the results provided herein focus on the sented an approximate equation for practical bridge bearings with
effective linear stiffness, accounting for softening that occurs at large shape factors, where the a term would be 6k and k is related
moderate strain levels that exceed typical service limits. to rubber hardness and can be estimated between 0.75 and 0.55 for
The compression and shear stiffness results from the testing 50 to 65 Shore A hardness elastomers. In the AASHTO (2008a)
program are summarized in Table 3, Kv represents the vertical specifications, the commentary provides this equation with a 5 6,
stiffness of the bearing determined from the experimental load- although a ballot has been proposed to modify the value from 6 to 4.8
deflection curve, as shown in Fig. 6. The values are linear regressions (Daniel H. Tobias, personal communication, 2012).
of the experimental data for a region bounded by the first instance Kelly and Konstantinidis (2011) provide extensive treatment of
when the loading curve reached 50% of the target vertical load and compression stiffness for various geometric arrangements and
the end of the unloading/reloading cycle. Although the loading rate boundary conditions. If the relaxation permitted by friction is
used to determine Kv was roughly 20 times faster than that specified neglected, a for the bearings in this study would be 6.37 and 6.50 for
in AASHTO (2008b) M251 and ASTM (2007) D4014, an exper- the small and large bearings, respectively. If the lack of restraint
imental study of load rate sensitivity showed that the increase in provided by friction only at the bottom surface of the bearing is
vertical stiffness caused by the faster loading rate was only about 5% considered in accordance with Kelly and Konstantinidis (2009), and
when compared with the AASHTO (2008b) and ASTM (2007) the bearings are assumed to be infinite strips with the narrow di-
loading protocols. The reported Kv values are also larger than values mension used as the width of the strip, the vertical stiffness decreases
that would be obtained from AASHTO (2008b) or ASTM (2007) by 4.9% if the coefficient of friction, m, between the rubber and
procedures because the range of data used in the regression was concrete is as high as 0.5, or up to 13.0% if m is as low as 0.2.
restricted to a stiffer portion of the loading curve. Consequently, Reductions in a values were calculated for each specimen,

JOURNAL OF BRIDGE ENGINEERING © ASCE / JULY 2013 / 617

J. Bridge Eng., 2013, 18(7): 612-623


Table 3. Vertical Stiffness and Apparent Shear Moduli
Test number Kv (kN/mm) GEst,Kv (kPa) GLin (kPa) G25 (kPa) G50 (kPa) g max,test (%) g max,specimen (%)
1-1 136 593 565 962 678 220 220
1-2 134 583 465 700 553 195 220
1-3 134 583 436 631 533 206 220
1-4 143 623 513 534 494 206 220
1-5 138 601 423 524 489 185 220
1-6 97 424 390 544 484 136 220
1-7 133 581 452 493 466 206 220
1-8 134 586 582 730 617 229 229
Downloaded from ascelibrary.org by CRRI - Central Road Research Institute on 01/09/18. Copyright ASCE. For personal use only; all rights reserved.

1-9 135 590 427 471 444 210 229


1-10 128 557 416 646 512 227 229
2-1 110 461 Varies Varies Varies Varies 127
2-2 152 673 Varies Varies Varies Varies 244
3-1 132 575 Varies Varies Varies Varies 211
3-2 122 533 Varies Varies Varies Varies 224
4-1 304 593 Varies Varies Varies Varies 152
4-2 292 588 497 882 736 135 152
4-3 288 581 626 972 802 158 158

Fig. 6. Determination of vertical stiffness

reflecting the frictional restraint observed during sliding in the re- Calculation of effective linear stiffness required identification
spective experiments, with a final value calculated for each specimen of slip initiation. To identify slip initiation, data from the sensors
by considering the individual layers of rubber as a series of springs measuring displacements at the two welded steel studs (dh,high in
(Kelly and Konstantinidis 2011). The resulting range of a factors Fig. 7) were compared with the average displacement measurements
was 5.80–6.02 for the small bearings and 5.92–6.10 for the large of four points near the base of the elastomer (dh,low in Fig. 7). Shear
bearing. Applying the relationship between compression and shear strain was estimated according to
stiffness, an estimate of shear modulus from a compression test is
obtained by substituting ɛ 5 dv =ERT, s 5 Fv =A, and Kv 5 Fv =dv , to  
dh,high 2 dh,low
obtain gshear ¼ P (3)
h 2 tshim,incl
Kv hrt
GEst,Kv ¼ (2)
aAS2 where h 5 height from the lower P attachment points to the un-
derside of the top steel plate, and tshim,incl 5 total thickness of the
In addition to this estimated material stiffness parameter, several shims included in h. Slip corresponds to regions of measured
apparent shear moduli were also calculated from the experimental response with incremental Dgshear  0. Points identified as slip
results. initiations were used to bound ranges for linear regression of shear

618 / JOURNAL OF BRIDGE ENGINEERING © ASCE / JULY 2013

J. Bridge Eng., 2013, 18(7): 612-623


stiffness to the experimental results, and secant stiffnesses were The results for the apparent shear moduli according to the various
also calculated to bounds of 25 and 50% ESS beyond the first data approaches are summarized in Table 3 for monotonic tests, sup-
point at which the horizontal load crossed zero in the direction of plemented with Fig. 8 for cyclic tests.
loading. These fixed ranges were chosen to approximately reflect Fig. 8 illustrates that apparent shear moduli degrade with in-
stiffness estimations consistent with ASTM 4014 and AASHTO creasing shear strain, and the data also suggest that the apparent
M251, respectively. The 25 and 50% secant stiffnesses and the reduction in stiffness at large strain levels is influenced by other
slope of the linear fit are converted to apparent shear moduli ac- factors besides material scragging. For example, Tests 2-1 and 2-2
cording to show that the apparent shear modulus at low strain levels in the
second test is reflective of the previously attained 125% strain level
DFh hrt in the first test, and only recovers partially in the interim of several
G¼ (4)
Ddh A days between the tests, when the bearing was dismounted from the
Downloaded from ascelibrary.org by CRRI - Central Road Research Institute on 01/09/18. Copyright ASCE. For personal use only; all rights reserved.

test frame and therefore completely unloaded. The trailing edge of


the base of the elastomer was observed to curl, and the leading edge
The apparent shear moduli corresponding to 25 and 50% shear
was rounded off by being ground against the concrete. These effects
strain, denoted as G25 and G50 , respectively, provide context for
contributed to the observed stiffness degradation through reduced
comparison with recognized standards and illustrate the general
mechanical and geometric participation of the elastomer. The re-
trend of shear constitutive response for increasing shear demands.
duction as a result of the concentrated abrasion at the leading edge
However, GLin , determined from the shear stiffness bounded by slip
escalated with increasing numbers of slip cycles such that the
initiation, captures the reduced stiffness associated with strains
maximum degradation from the initial stiffness of the first 25%
exceeding typical service-level limits, and is therefore a preferred
quarter cycle reached about 40–50%, compared with about 15–25%
parameter for evaluating the performance of quasi-isolated systems.
reductions through the 50% peak strain cycles. Additionally, when
the bearing was subjected to deformation at a high strain rate, the
observed stiffness was larger than for a low strain rate, as illustrated
by Tests 4-1 and 4-3.
When combined with the friction slip force limits, the apparent
shear modulus provides insight into the dynamic characteristics of
a bridge system prior to slip and the amount of deformation that can
be sustained without incurring a residual offset in bridge position
following an earthquake. However, care should be taken to ap-
propriately adjust the values obtained from standard tests to reflect
the various influences on sliding bearing response. The secant
stiffness used to characterize shear response up to about 50 to 65%
tended to overestimate dynamic, larger strain shear stiffness, es-
pecially with increasing slip cycles in Test 3-2. Alternatively, shear
stiffness estimated from compression stiffness overestimated the
Fig. 7. Measurement of shear deformation values observed during Test 3-2 but underestimated shear stiffness
for Test 4-3.

Fig. 8. Linear shear moduli for cyclic tests

JOURNAL OF BRIDGE ENGINEERING © ASCE / JULY 2013 / 619

J. Bridge Eng., 2013, 18(7): 612-623


Slip Response of Elastomeric Bearings on Concrete However, for a quasi-isolated system the influence of the true friction
value is more significant than in design scenarios to prevent walking,
Friction resistance has been studied for various combinations of
and the conservatism normally exercised in design considerations
natural rubber and neoprene elastomeric materials in plain and
to prevent walking could have other implications for global bridge
laminated configurations with steel, smooth concrete, and rough
performance in seismic scenarios. As the coefficient of friction
concrete mating surfaces and at various compression levels and
increases, the peak and residual seismic displacements of the su-
shearing velocities, producing a broad range of potential friction
perstructure diminish, while forces transmitted down into the
thresholds (Schrage 1981; Muscarella and Yura 1995; McDonald
substructure may increase. Therefore, considerable effort has been
et al. 2000). The bearings in the current testing program are full-scale
expended to characterize the expected slip behavior for large
laminated elastomeric bearings, using natural rubber, and as in-
deformations and multiple cycles.
dicated in Table 2, the testing program included multiple com-
As has been observed in previous tests, there is a significant
pression levels and shearing velocities. Characterization of friction
Downloaded from ascelibrary.org by CRRI - Central Road Research Institute on 01/09/18. Copyright ASCE. For personal use only; all rights reserved.

correlation between the friction coefficient and normal stress at the


mating surfaces is often subjective, using terms such as rough or
slip interface. Schrage (1981) observed an inversely proportional
smooth. For this testing program, the concrete surfaces were
trend of the friction coefficient versus compression stress. The data
roughened following IDOT (2007) standard specifications; how-
obtained from the current testing program agree with this general
ever, the guidance in that case is also subjective, calling for a uni-
trend, as seen in Fig. 9 (these data were averaged over 25% ESS slip
form, slightly roughened surface. To correlate the surface roughness
segments for each plotted point in both Figs. 9 and 10). In the QS
to a recognized measurement system, the surfaces were visually
tests, the data are clustered around the target vertical load on the
classified as having a rating of 3 or 4 using concrete surface profile
horizontal axis of the plot. During the ISR tests, there was a wider
chips produced by the International Concrete Repair Institute
variation of imposed vertical load, with the average compression
(ICRI 1997).
stress ranging from 1.59 to 5.30 MPa (230 to 770 psi), arising from
A primary emphasis in previous studies has been to establish
differing hydraulic capacities and demands among the actuators.
a reliable threshold to prevent slip and walking during service
Starting from the form proposed by Schrage (1981), for the cyclic
loading, where walking is the phenomenon of bearings gradually
quasi-static tests, the residual of the observed and expected
accruing sliding displacements on substructures over extended
coefficients of friction can be accounted for approximately
periods of time. Such a phenomenon was clearly seen at an accel-
according to
erated rate in this testing program during pauses in the tests (with the P
top plate held in place by the test frame while the base of the elas- 0:37 duslip
mQS ¼ 0:18 þ 2 (5)
tomer was only restrained by friction) that coincided with high s 100
elastomeric shear strains. The base of the bearing continued to slip P
during such intervals, with observable reduction in the horizontal where duslip 5 total aggregation of slip travel for the bearing (m),
force required to hold the top plate in position, reflecting the offset of and s 5 average compression stress acting on the bearing (MPa).
walking slip with reduced elastomer shear strains. The variation of the coefficient of friction with average compression
To address slip in service, the AASHTO (2008a) specifications stress and total slip is less amenable to fitting with minor adjustments
recommend an assumed design coefficient of friction of 0.2 between to Schrage’s original formula. However, those data are only rep-
the elastomer and concrete. The 0.2 value is presumed to be con- resentative of a single cyclic test and two monotonic tests, thus there
servative in the sense of underestimating the true resistance to slip. is not yet sufficient data to conclusively correlate the discrepancy in

Fig. 9. Observed friction versus average vertical compression stress

620 / JOURNAL OF BRIDGE ENGINEERING © ASCE / JULY 2013

J. Bridge Eng., 2013, 18(7): 612-623


Downloaded from ascelibrary.org by CRRI - Central Road Research Institute on 01/09/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 10. Observed friction versus total slip travel

the observed trend with respect to Schrage (1981) to the slip rate. The degradation as a function of cumulative slip in Test 2-2 was similar
ISR tests displayed marked increases in breakaway friction coef- to Test 2-1.
ficients over the QS test cases. For Specimen 1, the friction co- To put these cumulative slip values into perspective, the total
efficient rose from an average of 0.29 in QS Test 1-7 to an initial slip travel demands estimated from system analyses of full bridge
breakaway coefficient of 0.42 in ISR Test 1-8. Similarly, after models subjected to 2,500-year return period earthquake motions
Specimen 4 had exhibited sliding at an average friction coefficient for southern Illinois were 2.15 m (85 in.) (Filipov et al. 2011),
of 0.27 for the final slip excursion of QS Test 4-1, and 0.30 for QS compared with the total slip travel demands of about 4 m (160 in.)
Test 4-2, the initial breakaway coefficient rose to 0.46 in ISR Test imposed on the bearing specimens during the tests. When the
4-3. The friction coefficients obtained from the QS tests appear to be expected total slip travel during major earthquakes is taken into
consistent with the kinetic slip coefficients in the ISR tests. account, the data from these tests suggest bearings that were not
Similar increases in the static friction breakaway coefficient designed to tolerate seismic demands can be expected to survive
relative to the kinetic friction coefficient have been observed for large slip travel demands with minor visible abrasions, while
tests of rubber on concrete, with a kinetic friction coefficient of 0.8 maintaining reliable strength to support vertical loads and pro-
and a static friction coefficient of at least 1 (Serway and Beichner viding stable hysteretic shear sliding response to dissipate seismic
2000). Friction tends to reach a peak and then drop off at higher energy.
velocities, where the velocity at which the peak occurs is a function
of the hardness of the mating surfaces and the size of the surface
asperities (Persson 2001). Recent tests reported by Guo et al. (2006) Summary and Conclusions
indicate that maximum friction coefficients tend to occur around slip
velocities of 1 m/s for rubber tires on dry concrete. Therefore, for Tests were performed to characterize the elastomeric shear de-
large earthquakes the coefficient of friction could be higher than the formation and slip response of steel-reinforced, laminated bearings
values observed in this testing program. unrestrained against horizontal translation atop concrete sub-
For the QS tests, the observed friction coefficient degraded with structures. All bearings were composed of steel-reinforced natural
increasing cycles. For example, during Test 4-1 Fig. 9 shows the rubber, vulcanized to top steel plates. The tests included various
coefficient of friction varying from a maximum of about 0.45 to combinations of bearing size, vertical stress, and shearing rate, with
a minimum of about 0.25, and Fig. 10 shows that the general trend is monotonic and fully reversed cyclic loading protocols. Two vertical
to degrade with increasing cumulative slip imposed on the bearing. actuators and a horizontal actuator were used to impose target hor-
Generally, visible damage was limited to chamfered corners and izontal displacements in a range of 6 400% of total rubber thickness,
abrasions on the base of the bearing, with portions of elastomer that while maintaining an approximately constant vertical load. The
had been ground off remaining on the concrete surface after com- bearing specimens reacted at their base against concrete test pads
pletion of the tests. The most severe damage was observed in with roughened surfaces conforming to standard specifications
Specimen 2. That specimen was first subjected to about 3.05 m used by IDOT.
(120 in.) of total slip travel at a normal force of approximately 75 kN The experimental program has produced test data indicating that
(16.8 kip) during Test 2-1. The bearing was then further subjected to typical bearings used by IDOT can generally be characterized with
an additional slip travel of about 1.02 m (40 in.) with a normal force linear elastic response for low-to-moderate loads, even at large
of approximately 300 kN (67.2 kip). At the end of Test 2-2, abrasion displacement demands on the order of 250% shear strain. Addi-
of the elastomer base edges was sufficient to expose the edge of the tionally, the elastomer compound demonstrated appreciable damp-
lowest internal shim. Despite this damage, the rate of slip coefficient ing as an unintentional consequence of the measures taken to

JOURNAL OF BRIDGE ENGINEERING © ASCE / JULY 2013 / 621

J. Bridge Eng., 2013, 18(7): 612-623


provide hardness sufficient to meet standard specifications used by FHwA. The authors would like to thank the members of the project
IDOT. Development of a pronounced stiffening branch is precluded Technical Review Panel, chaired by D. H. Tobias of the Illinois
because the elastomer is permitted to slide on the concrete sub- Department of Transportation, for their valuable assistance with this
structure, although the force-displacement response near slip may research.
stiffen slightly for large gravity load stresses. The shear stiffness
observed at moderate strains showed significant degradation, with
References
reductions up to about 40–50% from the initial shear stiffness ob-
served at 25% shear strain. Either values taken from standard shear
AASHTO. (2008a). AASHTO LRFD bridge design specifications,
stiffness tests or estimated values derived from compression stiff-
AASHTO, Washington, DC.
ness provide some insight into the stiffness to be expected at strains AASHTO. (2008b). “Standard specification for plain and laminated elasto-
exceeding service limits; however, neither is perfectly suited to meric bridge bearings.” AASHTO M251-06, Washington, DC.
Downloaded from ascelibrary.org by CRRI - Central Road Research Institute on 01/09/18. Copyright ASCE. For personal use only; all rights reserved.

defining a reliable value for dynamic shear stiffness at strain levels AASHTO. (2010). Guide specifications for seismic isolation design,
expected during large earthquakes. AASHTO, Washington, DC.
Repeated monotonic tests on a single bearing, along with cyclic ASTM. (2007). “Standard specification for plain and steel-laminated
tests incorporating multiple reversed cycles up to 400% equivalent elastomeric bearings for bridges.” D4014-03, West Conshohocken, PA.
shear strain, indicate that the bearings are remarkably resilient. Buckle, I. G., and Mayes, R. L. (1990). “Seismic isolation: History, ap-
Scragging effects on stiffness appear minor in relation to the more plication, and performance—A world view.” Earthquake Spectra, 6(2),
significant influence of maximum shear strain demand. The bottom 161–201.
Buckle, I., Nagarajaiah, S., and Ferrell, K. (2002). “Stability of elasto-
surface of the elastomer is visibly worn and roughened by abrasion
meric isolation bearings: Experimental study.” J. Struct. Eng., 128(1),
with the concrete through multiple cycles of large displacements; 3–11.
however, this visible damage appears to have only a minor influence Calvi, G. M., and Pavese, A. (1997). “Conceptual design of isolation systems
on the force-displacement response. Additionally, delamination for bridge structures.” J. Earthquake Eng., 1(1), 193–218.
was not observed externally, and the only visible damage extend- Filipov, E. T., et al. (2013a). “Seismic performance of highway bridges
ing to the internal shims was for the case of the highest tested with fusing bearing components for quasi-isolation.” Earthquake En-
vertical load. gineering and Structural Dynamics, in press.
The bearing slip response gradually transitioned through a range Filipov, E. T., Fahnestock, L. A., Steelman, J. S., Hajjar, J. F., LaFave, J. M.,
of shear strains, from shear dominant to slip dominant. Therefore, and Foutch, D. A. (2013b). “Evaluation of quasi-isolated seismic bridge
even when displacements arise primarily from shear deformations, behavior using nonlinear bearing models.” Eng. Struct., 49(14), 168–
181.
a bearing may experience an accelerated walking phenomenon
Filipov, E. T., Hajjar, J. F., Steelman, J. S., Fahnestock, L. A., LaFave, J. M.,
for long duration earthquakes. The observed slip coefficients varied and Foutch, D. A. (2011). “Computational analyses of quasi-isolated
significantly with vertical load, following a nonlinear inverse re- bridges with fusing bearing components.” Proc., ASCE/SEI Structures
lationship. Deterioration of slip resistance is also noticeable over Congress, ASCE, Reston, VA, 276–288.
significant accumulations of slip for a constant vertical load. Guo, K.-h., Zhuang, Y., Chen, S.-k., and William, L. (2006). “Experimental
The stiffness and slip characteristics obtained from this series of research on friction of vehicle tire rubber.” Front. Mech. Eng. China,
experiments have been incorporated into full bridge system analyses 1(1), 14–20.
to evaluate the global response resulting from the use of bearings Highway Innovative Technology Evaluation Center (HITEC). (1996).
such as those in this paper (Filipov et al. 2013a; Filipov et al. 2013b). Guidelines for the testing of seismic isolation and energy dissipating
The global system analyses provide a comprehensive estimation of devices, ASCE, Reston, VA.
Illinois Department of Transportation (IDOT). (2007). Standard specifi-
seismic peak and residual displacements, accounting for the in-
cation for road and bridge construction, IDOT, Springfield, IL.
fluence of the abutments in the longitudinal direction, the stiffened Illinois Department of Transportation (IDOT). (2009). Bridge manual,
angle retainer brackets anchored to the concrete substructure in the IDOT, Springfield, IL.
transverse direction, as well as the bearing performance charac- International Concrete Repair Institute (ICRI). (1997). “Concrete surface
teristics observed from these experiments. Supplementary to the profile chips.” Æhttp://www.icri.org//bookstore/launchCatalog.asp?
experiments reported in this paper and the global bridge system ItemID=PCæ (Apr. 14, 2010).
analyses, additional tests incorporated in the full scope of the testing Kelly, J. M., and Konstantinidis, D. (2009). “Effect of friction on unbonded
program will provide further information regarding how the bearings elastomeric bearings.” J. Eng. Mech., 135(9), 953–960.
described in this paper behave under alternate situations and con- Kelly, J. M., and Konstantinidis, D. A. (2011). Mechanics of rubber
figurations, such as when transverse retainers are included, and bearings for seismic and vibration isolation, Wiley, Chichester, U.K.
Kikuchi, M., and Aiken, I. D. (1997). “An analytical hysteresis model for
together with three-dimensional systems models will offer insight
elastomeric seismic isolation bearings.” Earthquake Eng. Struct. Dyn.,
into how a quasi-isolated system performs during large magnitude 26(2), 215–231.
earthquakes. Konstantinidis, D., Kelly, J. M., and Makris, N. (2008). “Experimental
investigations on the seismic response of bridge bearings.” Rep. No.
EERC 2008-02, Earthquake Engineering Research Center, College of
Acknowledgments Engineering, Univ. of California at Berkeley, Berkeley, CA.
Kulak, R. F., and Hughes, T. H. (1992). “Mechanical tests for validation of
This paper is based on the results of ICT R27-70, Calibration and seismic isolation elastomer constitutive models.” Proc., Pressure Ves-
Refinement of Illinois’ Earthquake Resisting System Bridge Design sels and Piping Conf. on DOE Facilities Programs, Systems Interaction,
Methodology (LaFave et al. 2013). ICT R27-70 was conducted in and Active/Inactive Damping, C.-W. Lin and W. W. Chen, eds., Vol.
229, ASME, New York, 41–46.
cooperation with the Illinois Center for Transportation (ICT), IDOT,
LaFave, J., et al. (2013). “Experimental investigation of the seismic response
Division of Highways, and the U.S. Department of Transportation, of bridge bearings.” Research Report No. ICT-13-002, Illinois Center for
Federal Highway Administration (FHwA). The content of this paper Transportation, Univ. of Illinois at Urbana-Champaign, Urbana, IL.
reflects the view of the authors, who are responsible for the facts and Malaysian Rubber Producers’ Research Association (MRPRA). (1980).
the accuracy of the data presented herein. The content does not “Natural rubber engineering data sheet.” EDS 16, Malaysian Rubber
necessarily reflect the official views or policies of the ICT, IDOT, or Research and Development Board Organization, Hertford, U.K.

622 / JOURNAL OF BRIDGE ENGINEERING © ASCE / JULY 2013

J. Bridge Eng., 2013, 18(7): 612-623


Mayes, R. L., and Naeim, F. (2001). “Design of structures with seismic Schrage, I. (1981). “Anchoring of bearings by friction.” Special Publication
isolation.” The seismic design handbook, F. Naeim , ed., Kluwer, Boston, SP70-12, American Concrete Institute, Detroit, 197–215.
724–755. Serway, A. R., and Beichner, R. J. (2000). Physics for scientists and
McDonald, J., Heymsfield, E., and Avent, R. R. (2000). “Slippage of engineers, Saunders College Publishing, Philadelphia.
neoprene bridge bearings.” J. Bridge Eng., 5(3), 216–223. Shenton, H. W. (1996). “Guidelines for pre-qualification, prototype and
Mori, A., Moss, P. J., Cooke, N., and Carr, A. J. (1999). “The behavior of quality control testing of seismic isolation systems.” NISTIR 5800, NIST,
bearings used for seismic isolation under shear and axial load.” Gaithersburg, MD.
Earthquake Spectra, 15(2), 199–224. Stanton, J. F., and Roeder, C. W. (1982). “Elastomeric bearings design,
Muscarella, J. V., and Yura, J. A. (1995). “An experimental study of construction, and materials.” National Cooperative Highway Research
elastomeric bridge bearings with design recommendations.” Research Program 248, Transportation Research Board, National Research
Rep. No. 1304-3, Center for Transportation Research, Bureau of En- Council, Washington, DC.
gineering Research, Univ. of Texas at Austin, Austin, TX. Taylor, A. W., and Igusa, T. (2004). Primer on seismic isolation, ASCE,
Persson, B. N. J. (2001). “Theory of rubber friction and contact mechanics.”
Downloaded from ascelibrary.org by CRRI - Central Road Research Institute on 01/09/18. Copyright ASCE. For personal use only; all rights reserved.

Reston, VA.
J. Chem. Phys., 115(8), 3840–3861. Tobias, D. H., Anderson, R. E., Hodel, C. E., Kramer, W. M., Wahab, R. M.,
Roeder, C. W., Stanton, J. F., and Taylor, A. W. (1987). “Performance of and Chaput, R. J. (2008). “Overview of earthquake resisting system
elastomeric bearings.” National Cooperative Highway Research Program design and retrofit strategy for bridges in Illinois.” Pract. Period. Struct.
298, Transportation Research Board, Washington, DC. Des. Constr., 13(3), 147–158.

JOURNAL OF BRIDGE ENGINEERING © ASCE / JULY 2013 / 623

J. Bridge Eng., 2013, 18(7): 612-623

Anda mungkin juga menyukai