Sie sind auf Seite 1von 25

Bulletin of Earthquake Engineering (2005) 3:7599

DOI 10.1007/s10518-005-0187-9

Springer 2005

Frictional Behavior of Steel-PTFE Interfaces


for Seismic Isolation
M. DOLCE , D. CARDONE and F. CROATTO
DiSGG, University of Basilicata, Macchia Romana Campus, 85100, Potenza, Italy

Corresponding author. Tel: +390971205052; Fax: +390971205052, E-mail:


dolcerom@libero.it
Received 17 October 2004; accepted 6 December 2004
Abstract. The widespread use of sliding bearings for the seismic isolation of structures
requires detailed knowledge of their behavior and improved modeling capability under seismic conditions. The paper summarizes the results of a large experimental investigation on
steelPTFE interfaces, aimed at evaluating the effects of sliding velocity, contact pressure,
air temperature and state of lubrication on the mechanical behavior of steel-PTFE sliding
bearings. Based on the experimental outcomes, two different mathematical models have been
calibrated, which are capable of accounting for the investigated parameters in the evaluation of the sliding friction coefcient. The rst model is basically an extension of the model
proposed by Constantinou et al. (1990) Journal of Earthquake Engineering, 116(2), 455472,
while the second model is derived from the one proposed by Chang et al. (1990) Journal of
Engineering Mechanics, 116, 27492763. Expressions of the model parameters as a function
of bearing pressure and air temperature are presented for lubricated and non-lubricated sliding surfaces. Predicted and experimental results are nally compared.
Key words: friction, PTFEsteel interfaces, seismic isolation, sliding bearings

1. Introduction
SteelPTFE sliding bearings have been widely used in the past to accommodate thermal movements and effects of pre-stressing, creep and shrinkage in bridges. More recently, they have been proposed as part of seismic
isolation systems, to support the weight of the structure, while relying upon
separate mechanisms, to provide the system with re-centring and additional
energy absorbing capability.
Several sliding isolation systems with restoring force have been proposed. Some of them have reached the stage of implementation, such as the
Resilient-Friction Base Isolation System (Mostaghel, 1984), based on the
elastic properties of rubber, the SMA Isolation System (Dolce et al., 2000),
based on the superelastic properties of shape memory alloys, and the Friction Pendulum System (Zayas et al., 1987), which exploits an articulated
slider moving on a spherical surface to provide restoring capability. The
most important advantage in using sliding isolation systems with (weak)

76

M. DOLCE ET AL.

restoring force is that the structural response is little sensitive to variations


in the frequency content of the earthquake.
In Italy a different approach has been followed (Dolce, 2001). A very
large number of bridges, indeed, have been equipped with elasto-plastic isolation systems, consisting of lubricated sliding bearings and hysteretic steel
dampers. Such systems are able to limit the force transmitted by the deck
to piers and abutments to a predened level, almost independent from the
intensity and spectral content of the seismic excitation. The drawbacks are
the large dispersion in the peak displacements and the occurrence of permanent displacements after strong earthquakes.
Many numerical analyses of the seismic response of structures equipped
with steelPTFE sliding isolation systems have been carried out (Mostaghel
and Tanbakuchi, 1983; Fan et al., 1988). In all these studies, as well as
in the current design practice, it is assumed that the friction coefcient
comply with the Coulomb friction law (i.e. friction remains constant during sliding). Actually, experimental observations (Constantinou et al., 1987;
Hwang et al., 1990; Mokha et al., 1990), pointed out that the friction
coefcient increases more than linearly while increasing sliding velocity,
while it decreases with the increase of the contact pressure. Temperature
and number of sliding reversals also play a not negligible role (Tyler, 1977).
An accurate mathematical model of the frictional behavior of steel
PTFE sliding bearings has been developed in (Constantinou et al., 1990).
It is based on the viscoplasticity theory (Wen, 1976), and is referred to
as modied viscoplastic model. Its main characteristic is the dependence
of the friction coefcient on sliding velocity and bearing pressure, through
an exponential analytical law. The Constantinous model has been implemented in the structural analysis program SAP-2000 Nonlinear (SAP-2000,
2002), as the Friction-Pendulum Isolator NLLink element. Applications of
the exponential model in the analysis of a sliding isolation system have
been reported in (Constantinou et al., 1990; Mokha et al., 1993; Deb and
Paul, 2000), mainly with the scope of evaluating the effects of bearing pressure, sliding velocity, breakaway friction and bi-directional motion on the
seismic response of base-isolated buildings, compared with the predictions
of the Coulombs model. An interesting comparison between experimental
and numerical results is reported in (Tsopelas et al., 1996), with reference
to a bridge structure.
The Constantinous model is a phenomenological model, being derived
from the observation of experimental results. An evolution of the
Constantinous model, based on the tribology theory, has been recently
proposed in (Takahashi et al., 2004) to describe the frictional behavior of
PTFEsteel interfaces at the microscopic level.
A comprehensive program of experimental tests has been carried out at
the laboratory of the University of Basilicata on steelPTFE interfaces, in

FRICTIONAL BEHAVIOR OF STEELPTFE INTERFACE

77

order to fully investigate the effects of sliding velocity, contact pressure, air
temperature, number of cycles and state of lubrication on the mechanical
behavior of steelPTFE sliders. Based on the experimental outcomes, two
mathematical models of their frictional behavior, for conditions of interest
in seismic isolation, have been developed and calibrated. The rst model
is basically an extension of the model proposed in (Constantinou et al.,
1990), while the second model is derived from the one proposed in (Chang
et al., 1990). In this paper, the main results of the experimental tests are
described. Model predictions and experimental results are then compared.
2. Experimental Set up and Procedure
The experimental program on steelPTFE interfaces had two specic objectives: (i) investigating the variability of the sliding friction coefcient while
varying contact pressure, velocity, air temperature, displacement amplitude
and state of lubrication of steelPTFE interfaces, (ii) developing and calibrating a numerical model of the mechanical behavior of steelPTFE sliding bearings.
2.1. Materials
The materials used were as follows:
Unlled PTFE pads obtained from a 5.45 mm thick sheet with dimpled
recesses, whose dimensions and pattern are shown in Figure 1. The function of the dimple recesses is to retain the grease and to gradually introduce it between the sliding interfaces, during wear of PTFE.
Stainless steel plates (AISI 316/L) of 3 mm thickness, polished to mirror
nish, with less than 0.1 m surface roughness.
Silicone grease, of the type normally used in the bearing manufacturing.

Figure 1. Details of PTFE dimpled recesses.

78

M. DOLCE ET AL.

2.2. Test apparatus


The testing apparatus is schematically shown in Figure 2. A central steel
plate is laterally nished with two polished stainless steel plates (300 mm
by 45 mm dimensions) and sandwiched between two couples of 40 by
25 mm PTFE pads, with 5.45 mm thickness. The PTFE pads are placed
into recesses of two lateral steel plates. The protrusion of the pads was
2.6 mm in the unloaded condition.
The PTFEsteel interfaces are compressed by a 50 kN hydraulic jack
and four pre-stressing steel rods. The lateral steel plates are attached to a
reaction frame, while the central steel plate is driven by a 10 kN Instron
dynamic actuator, with 125 mm stroke. The dynamic actuator is equipped
with a 63 lit/s controller valve, a 10 kN load cell and a 125 mm internal inductive transducer. The above said test arrangement is placed inside a
thermal room, working in 30 C/+80 C temperature range. The air temperature is controlled by a K-type thermocouple.

Figure 2. Testing apparatus.

79

FRICTIONAL BEHAVIOR OF STEELPTFE INTERFACE

2.3. Test program


More than 300 tests have been carried out on steelPTFE interfaces with
9.36, 18.72 and 28.1 MPa bearing pressures, 10, 20 and 50 C air temperatures, from 1mm/s to about 300 mm/s sliding velocities, from 10 to 50 mm
displacement amplitudes. Both lubricated and non-lubricated steelPTFE
interfaces were tested. In most of the tests, the motion was sinusoidal, with
specied amplitude and frequency. Furthermore, a number of tests were
conducted with constant velocity motion (i.e. saw-tooth displacementtime
history). The two types of tests produced almost identical results for the
same peak velocity of sliding.
Table I summarises the whole experimental program. As can be seen,
nine series of tests have been carried out. Each series consisted of 16
tests, all at the same air temperature and contact pressure. The contact
pressure was progressively increased from one series to another, while keeping the temperature constant. Every three series of tests the PTFE pads
were replaced by new ones.

Table I. Test program.


Test Interfacesa T b ( C)
No.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
a

L/P
L/P
L/P
L/P
L/P
L/P
L/P
L/P
L/P
L/P
L/P
L/P
L/P
L/P
L/P
L/P

10/20/50
10/20/50
10/20/50
10/20/50
10/20/50
10/20/50
10/20/50
10/20/50
10/20/50
10/20/50
10/20/50
10/20/50
10/20/50
10/20/50
10/20/50
10/20/50

P c (MPa)

dd
fe
Wavef
(mm) (Hz)

vg
Cycles CDh
(mm/s) No.
(mm)

9.36/18.7/28.1
9.36/18.7/28.1
9.36/18.7/28.1
9.36/18.7/28.1
9.36/18.7/28.1
9.36/18.7/28.1
9.36/18.7/28.1
9.36/18.7/28.1
9.36/18.7/28.1
9.36/18.7/28.1
9.36/18.7/28.1
9.36/18.7/28.1
9.36/18.7/28.1
9.36/18.7/28.1
9.36/18.7/28.1
9.36/18.7/28.1

50
50
50
50
10
10
10
10
25
25
25
25
50
50
50
50

10
40
100
200
var.
var.
var.
var.
var.
var.
var.
var.
var.
var.
var.
var.

0.05
0.2
0.5
1
0.05
0.2
0.5
1
0.05
0.2
0.5
1
0.05
0.2
0.5
1

TR.
TR.
TR.
TR.
SYN.
SYN.
SYN.
SYN.
SYN.
SYN.
SYN.
SYN.
SYN.
SYN.
SYN.
SYN.

State of stainless steelPTFE interfaces : L = lubricated, P = Pure.


Air temperature.
c
Contact Pressure.
d
Displacement amplitude.
e
Frequency of loading.
f
Displacement prole: TR = triangular, SYN = sinusoidal.
g
Sliding velocity: var = variable.
h
Cumulative distance.
b

5
5
5
5
5
5
5
5
5
5
5
5
5
5
5
5

1000
2000
3000
4000
4200
4400
4600
4800
5300
5800
6300
6800
7800
8800
9800
10,800

80

M. DOLCE ET AL.

Triangular (saw-tooth) tests were carried out at constant displacement amplitude (50 mm), while increasing the sliding velocity from 10 to
200 mm/s. In the sinusoidal tests, the displacement amplitude was varied
from 10 to 50 mm, while increasing the frequency of loading from 0.05 to
1 Hz, thus producing peak sliding velocities ranging from 1 to 300 mm/s.
Five complete loading cycles were performed during each test. The cumulative distance during each series of test was approximately 10.8 m, as shown
in Table I. Thus, a total distance of about 32.4 m was covered by each set
of PTFE pads. The thickness of the PTFE pads was measured before testing and after their removal.
3. Experimental Results
3.1. Effect of type of test
Representative frictional force-displacement loops of the single interface are
shown in Figures 3 and 4. They refer to tests conducted on non-lubricated interfaces under the same contact pressure (28.1 MPa), air temperature (10 C) and displacement amplitude (50 mm), only differing for the
wave form and the loading frequency. Figure 3 compares the frictional
behavior exhibited by the PTFEsteel interfaces in (a) triangular and (b)
sinusoidal tests at the same peak velocity, equal to 10 mm/s (i.e. test No. 1
vs. test No. 13, according to Table I). Figure 4 refers to (a) triangular and
(b) sinusoidal tests conducted at the higher peak velocities: 200 mm/s and
300 mm/s, respectively (i.e. test No. 4 vs. test No. 16, according to Table I).
For the sinusoidal tests, the peak velocity is dened as the average velocity
in the displacement range corresponding to force levels greater than 95%
of the maximum frictional force.
Two phenomena are clearly visible in the triangular tests, one at the
start of sliding, the other at every motion reversal. The rst phenomenon
is generally taken into account through the denition of a breakaway friction coefcient, also known as static friction coefcient, to distinguish it
from the sliding (kinetic) friction coefcient. The second phenomenon, generally referred to as stick-slip, corresponds to a short duration increase of
the frictional force, followed by a rapid decrease. Both the observed experimental phenomena can be related to (i) a momentary sticking of the interfaces and to the (ii) acceleration impulse occurring at the start of the test
and at every motion reversal, especially in the triangular tests.
The examination of Figures 3 and 4 clearly highlights the dependence
of the friction coefcient from sliding velocity. In each triangular test, the
forcedisplacement loops are practically rectangular, in accordance with
friction Coulombs law, but the friction force increases while increasing the
frequency of loading in the different tests. In the sinusoidal tests, on the

81

FRICTIONAL BEHAVIOR OF STEELPTFE INTERFACE

(mm)

60
0
160
(sec)

(KN)

-60

s t ick slip

break-away friction

4
2
0
-60

-40

-20

20

40

-2

60
(mm)

-4
-6

(a)

stick slip

(KN)

(mm)

60
0
0

160
(sec)

-60

6
4
2
0

-60

-40

-20

0
-2

20

40

60
(mm)

-4
(b)

-6

Figure 3. Typical frictional forcedisplacement loops recorded during (a) triangular


and (b) sinusoidal tests at low peak velocities (about 10 mm/s). Both tests were conducted on non-lubricated interfaces, under the same air temperature (10 C), contact
pressure (28.1 MPa) and displacement amplitude (50 mm).

contrary, the friction coefcient varies during the motion, reaching its maximum at the maximum velocity (i.e. at zero displacement). Furthermore the
friction coefcient tends to decrease during continuous loading cycles, this
effect being related to self-heating of the sliding interfaces. Indeed, the rate
of decrease of the friction coefcient is quite negligible in the tests al low

82

M. DOLCE ET AL.

(mm)

60
0
20
(sec)

break-away friction

(KN)

-60

stick slip

8
6
4
2
0

-60

-40

-20

-2

20
4

-4

40

60
(mm)

-6
3

-8

stick slip

(a)

(KN)

(mm)

60
0
0

20

(sec)

-60

8
6
4
2
0

-60

-40

-20

-2
-4

20
4

40

60
(mm)

-6
(b)

-8

Figure 4. Typical frictional force-displacement loops recorded during (a) triangular


and (b) sinusoidal tests at very high peak velocities (>200 mm/s). Both tests were conducted on non-lubricated interfaces, under the same air temperature (10 C), contact
pressure (28.1 MPa) and displacement amplitude (50 mm).

frequency (see Figure 3), while it is decidedly more pronounced (>20% in


ve cycles) in the tests at high frequency of loading (see Figure 4). The
phenomenon comes to an end in a few cycles, due to the attainment of a
new thermal equilibrium with the ambient.

FRICTIONAL BEHAVIOR OF STEELPTFE INTERFACE

83

3.2. Effect of sliding velocity


Figure 5 summarizes the results of all the tests on non-lubricated interfaces,
in terms of sliding friction coefcient, i.e.:
=

Fr
N

(1)

where Fr is the frictional resistance of the sliding interfaces, and N is the


normal load applied during the tests, equal to 10, 20 and 30 kN, respectively.
In Figure 5, the sliding friction coefcient is reported as a function of
(peak) velocity and contact pressure, for three different temperature values:
(a) 10 C, (b) 20 C and (c) 50 C, respectively. In the calculation of the
sliding friction coefcient, reference was made to the second cycle of each
test, by averaging the values of the friction force associated to both motion
ways. In Figure 5, the experimental data are represented by single points,
while the curves correspond to analytical laws obtained from two different
frictional models, as discussed below.
As can be seen, the sliding friction coefcient increases rapidly with
velocity, up to a certain velocity value, beyond which it remains almost
constant. This value is around 150 mm/s, regardless air temperature and
bearing pressure.
The difference between maximum and minimum values of the sliding
friction coefcient (i.e.  = max min , see Figure 6(a)) is larger at low
contact pressures, being equal to about 12% at 9.36 MPa and less than 7%
at 28.1 MPa. The air temperature has little inuence on . On the contrary, the percent increment = /min (see Figure 6(b)) tends to increase
while increasing contact pressure, especially at medium-to-high temperatures, being of the order of 180% at 9.36 MPa, and 250% at 28.1 MPa.
In view of the use of steelPTFE sliding bearings in seismic isolation,
it is worth to remark that typical design values of displacement and frequency of vibration of isolated structures are between 100200 mm and
0.40.5 Hz, respectively. This means that the maximum sliding velocity,
occurring in steel-PTFE bearings during an earthquake, ranges between
160 mm/s and 400 mm/s. According to the experimental results, the sliding friction coefcient is practically constant for seismic applications, but
signicantly different from the friction coefcient in slow movements. For
the sliding interfaces considered in this study, the sliding friction coefcient
ranges between 10% and 11% (depending on temperature) for a bearing
pressure equal to 28.1 MPa, which is the closer experimental value to the
maximum allowable contact pressure under seismic condition (i.e. 41 MPa)
suggested in (AASHTO, 1999).

84

M. DOLCE ET AL.

P= 9.36 MPa

(%)

(a) 25

P= 18.72 MPa

P= 28.1 MPa

exponential

20
logarithmic

logarithmic

15
exponential

10
logarithmic

exponential

5
(mm/sec)

0
0
(%)

(b) 25

50

100

150

P= 9.36 MPa

200

250

P= 18.72 MPa

300

350

P= 28.1 MPa
exponential

20

logarithmic

logarithmic

15

exponential

10
logarithmic

5
exponential

(mm/sec)

(c) 25

50

(%)

100

150

P= 9.36 MPa

200

250

P= 18.72 MPa

300

350

P= 28.1 MPa
exponential

20

logarithmic

15

logarithmic
exponential

10
logarithmic

5
exponential

(mm/sec)

0
0

50

100

150

200

250

300

350

Figure 5. Variation of the friction coefcient with sliding velocity, air temperature
and bearing pressure, for non-lubricated interfaces. Comparison between analytical
laws and experimental results. Air temperature equal to: (a) 10 C, (b) 20 C and
(c) 50 C.

85

(%)

FRICTIONAL BEHAVIOR OF STEELPTFE INTERFACE


14

-10C

+20C

+50C

12
10
8
6
4
2
0
9.36

18.72

28.08
(MPa)

(%)

(a)

300

-10C

+20C

+50C

250
200
150
100
50
0
9.36
(b)

18.72

28.08
(MPa)

Figure 6. (a) Absolute and (b) percent increment of the sliding friction coefcient in
seismic with respect to service conditions, as a function of contact pressure, for three
different air temperatures.

During its service lifetime, the sliding isolator acts as a usual sliding
bearing, subjected to service load and thermal movements at very low
velocities. The operating conditions of sliding bearing contained in various codes (AASHTO, 1999), (BS 5400, 1983), (CEN 1337, 2000) are
quite different. CEN considers lubricated steelPTFE interfaces only, while
BS and AASHTO allow the use of both lubricated and non-lubricated
interfaces. The maximum allowable contact pressure is assumed equal to
24 MPa by AASHTO (in absence of specic wear tests), 45 MPa by BS
and 60 MPa by CEN. For unlled PTFE sliding against stainless steel without lubrication, the service friction coefcient suggested by AASHTO (in
absence of tests), for temperatures 25 C, is equal to 15% at 10 MPa
and 10% at 20 MPa (or more). At 20 C, instead, the service limit state friction coefcient is assumed equal to 6% at 10 MPa and 3% at 20 MPa (or

86

M. DOLCE ET AL.

more). It is worth to observe that the friction coefcient values obtained


in this experimental study, at the lowest sliding velocity (i.e. about 3 mm/s),
substantially agree with the above mentioned code limits, being equal to: (i)
8.9% at 9.36 MPa and 10 C, (ii) 6.6% at 18.7 MPa and 10 C, (iii) 6.5%
at 9.36 MPa and 20 C and (iv) 4.4% at 18.7 MPa and 20 C.
3.3. Effect of contact pressure
As known (Mokha et al., 1990), the sliding friction coefcient of steel
PTFE interfaces reduces while increasing pressure. Based on the experimental outcomes available (see Figure 5), the rate of reduction is practically
constant while increasing bearing pressure and quite insensitive to sliding
velocity and air temperature. As a matter of fact, by doubling the contact
pressure (from 9.36 to 18.7 MPa) the friction coefcient reduces, on average over the whole range of sliding velocities, by 24% (3.4%) at 10 C,
up to 33.4% (2.4%) at 50 C. Similarly, by tripling the contact pressure
(from 9.36 to 28.1 MPa) the friction coefcient reduces by 38.7% (3.6%)
at 10 C, up to 47.2% (3.9%) at 50 C. It is worth to remark that the
negative variation of the friction coefcient with contact pressure reduces
considerably the corresponding variation observed for the frictional force
when increasing the normal load.
3.4. Effect of air temperature
Figure 7 reports the friction coefcient at (a) very low (i.e. 8 mm/s) and
(b) very high (i.e. 316 mm/s) sliding velocities, as a function of air temperatures, for three different contact pressure values, namely: (i) 9.36 MPa, (ii)
18.7 MPa and (iii) 28.1 MPa. Experimental data (points) and model predictions (curves) are compared.
The analytical curves reported in Figure 7 refer to the logarithmic
model (see below), calibrated over the whole range of sliding velocities.
This explains some inconsistencies between experimental and numerical
data, especially at low sliding velocity.
Based on the experimental results, the sliding friction coefcient decreases
while increasing air temperature, according to a second-order polynomial
law (i.e. = a T 2 b T + c, with a, b, c > 0), whose expression is reported
in Figure 7, for 6 pressurevelocity couples of values. As can be noted, the
rate of reduction of the sliding friction coefcient is greater when passing
from low-to-medium temperatures than when passing from medium-to-high
temperatures. Moreover, it depends on sliding velocity, while being practically independent from contact pressure. At 8 mm/s (see Figure 7(a)), for
instance, the average rate of reduction of the sliding friction coefcient with
temperature is of the order of 0.77%/ C when passing from 10 to 20 C,

87

(%)

FRICTIONAL BEHAVIOR OF STEELPTFE INTERFACE

20

(P1 =9.36MPa, v1 =8mm/s): = 1.09E-1 - 7.31E-4*T + 4.3E-6*T2


(P2 =18.7MPa, v1 =8 mm/s): = 8.45E-2 - 7.31E-4*T + 4.3E-6*T2
(P3 =28.1MPa, v1 =8 mm/s): = 6.85E-2 - 7.31E-4*T + 4.3E-6*T2

15

in which T = (T-T0) and T0 = -10C


P1,v1

10

P2,v1

P3,v1

0
-10

(%)

(a)

20

20

P1,v2

15

P2,v2

50
(C)

P3,v2

10
(P1 =9.36MPa, v2 =316mm/s): = 1.96E-1 3.95E-4*T + 1E-6*T2
(P2 =18.7MPa, v2 =316mm/s): = 1.42E-1 3.95E-4*T + 1E-6*T2
(P3 =28.1MPa, v2 =316mm/s): = 1.18E-1 3.95E-4*T + 1E-6*T2

in which T = (T-T0) and T0 = -10C

0
(b)

-10

20

50
(C)

Figure 7. Sliding friction coefcient at (a) very low and (b) very high velocities (i.e. 8
and 316 mm/s, respectively) as a function of air temperatures, for three different normal pressure values (i.e. 9.36, 18.72 and 28.1 MPa, respectively). Comparison between
experimental results and model predictions.

while being of the order of 0.33%/ C when passing from 20 to 50 C. At


316 mm/s (see Figure 7(b)), the rates of reduction decrease by about 2.5
times.
To account for the changes in the friction resistance of steelPTFE sliding isolation systems due to temperature variations, the AASHTO denes
two system property modication factors, max,t and min,t , which quantify
the effects of temperature variations on the nominal value of the friction
coefcient at 20 C reference temperature. They are dened as the ratio of
the friction coefcient at the highest and at the lowest expected temperature (say 50 C and 10 C, respectively) to the friction coefcient at the
reference temperature (20 C). Specic tests lacking, AASHTO provides predetermined values: max,t = 1.2 and min,t = 1, for non-lubricated steel-PTFE
interfaces operating at 10 C and 50 C, respectively. The corresponding

88

M. DOLCE ET AL.

experimental values drawn from the whole set of experimental data are 1.17
(0.09) and 0.89 (0.054), respectively.
3.5. Effect of lubrication
Figures 8 and 9 show the sliding friction coefcient for a number of signicant tests on both non-lubricated (Figure 8) and lubricated (Figure 9)
interfaces. Each diagram of gure 8 and 9 refers to a different air temperature: (a) 10 C, (b) 20 C and (c) 50 C, respectively, as well as to a
different set of PTFE pads. Each series of data, instead, refer to a different
normal load, resulting in 9.36, 18.7 and 28.1 Mpa contact pressure, respectively. In each diagram the experimental data have been gathered in four
groups, based on the peak sliding velocity. Each group, therefore, corresponds to tests repeated under similar test conditions, after a certain number of cycles. The displacement amplitude is generally different from one
test to another, within each group. The order of execution of the tests is
given by the number beside each point.
By comparing Figures 8 and 9, it turns out that lubrication reduces friction coefcient by about 5 times at 10 and 20 C, and almost 8 times
at 50 C. The sliding friction coefcient of lubricated steelPTFE interfaces
results always below 4%, which is the limit value for sliding isolation devices
prescribed by the new Italian seismic code (Ordinanza 3274, 2003). The scatter in each group is practically negligible for non-lubricated interfaces, while
it is not for lubricated steelPTFE interfaces. For these latter, the tendency is
a progressive reduction of the friction coefcient while increasing the number of cycles performed, due to the following reasons: (i) the introduction,
in the sliding interfaces, of the grease contained inside the dimpled recessed,
during the wear of the PTFE pads and (ii) the test sequence (see Table I),
where tests wiith increasing displacement amplitude were carried out. Actually, before the start of each test sequence, grease was spread over the PTFE
pads, to ll the dimpled recesses, but a certain amount of grease was also
spread over the stainless steel sheets. When increasing displacement amplitude, it is reasonable to believe that more grease, from the stainless steel
sheets, was introduced in the sliding interfaces.
The quite negligible scatter in the experimental results relevant to nonlubricated interfaces (see Figure 8) implies that the wear of PTFE (see
below) do not affect signicantly their frictional behavior. For the lubricated steelPTFE interfaces, friction is much lower and the wear of PTFE
is expected to be much less than for non-lubricated interfaces.
Figure 10 summarizes the results of all the tests on lubricated PTFE
steel interfaces. The sliding friction coefcient is reported as a function of
(peak) velocity and contact pressure, for three different temperature values:
(a) 10 C, (b) 20 C and (c) 50 C, respectively. In the gure, experimental

89

FRICTIONAL BEHAVIOR OF STEELPTFE INTERFACE


10KN

(%)

20

15
6

10

17

22

20KN

30KN
8

11

24

27

40

43

10

13
18

23

26

34

39

42

15

28

31

44

47

30

29
45

12

14

46

33

38

50

10

50

50

10

25

10

25

50

25

50

10

12

15

40

32

30

67

77

60

155

152 (mm/sec)

20

15

10

2
1
17

13

22

29

18

10

23

26

34

39

42

45

(mm)

30KN

20KN

10KN

(%)

(a)

11

14

12
28

24

27

30

40

43

46

44

15
31

47

33

38

50

10

50

50

10

25

10

25

50

25

50

10

12

15

40

32

30

68

79

61

160

157 (mm/sec)

20

10KN

(%)

(b)

20KN

15
2

10

13

17

22

29

33

38

50
10

18

7
23

(mm)

30KN

11

24

27

12

15

28

31

44

47

14

10
26
40

43

30
46

45

34

39

42

10

50

50

10

25

10

25

50

25

50

12

15

40

34

31

70

80

60

162

159 (mm/sec)

(c)

(mm)

Figure 8. Effects of number of cycles on the sliding friction coefcient exhibited by


non-lubricated interfaces during tests at different air temperatures, namely: (a) 10 C,
(b) 20 C and (c) 50 C.

data and model predictions are compared. As can be seen, the friction
coefcient of lubricated interfaces follows a trend similar to that described
for non-lubricated interfaces. It tends to increase when increasing sliding

90

M. DOLCE ET AL.
10KN

(%)

20KN

38
33
17
1

42

18 23

45
13

50

50

46

24

34
29

22

44

43

39

30KN
40

25

10

31

12

11

10

28

30

27

26

47

15

14

0
50

(a)

10

12

15

31

31

10KN

64

50

79

25

60

17
33
1

22
38

29

20KN

160

30KN
11

13

40

25

50

(mm)

156

(mm/sec)

(%)

10

10

7
18
2

45

10

23
39

24

26
42

15

14

28

27

40

12

30

44

31

43
46

34

47

50

10

50

50

10

25

10

25

50

25

50

(mm)

10

12

15

40

32

30

65

80

61

159

154

(mm/sec)

20KN

10KN

(%)

(b)

30KN

1
17

6
22

2
13

29

39
18
23

38
45

33

8
10
42
26

40
24

11

43 14

44
12

15

28

47

46

27

31

30

34

(c)

50

10

50

50

10

25

10

25

50

25

50

(mm)

10

15

40

31

31

66

80

62

161

158

(mm/sec)

Figure 9. Effects of number of cycles on the sliding friction coefcient exhibited by


lubricated interfaces during tests at different air temperatures, namely: (a) 10 C, (b)
20 C and (c) 50 C.

velocity and to reduce when increasing air temperature and contact pressure. However a greater scatter can be noted, probably due to the testing
sequence and the difculty of getting uniform distribution of lubricating
grease. The behavior of the friction coefcient at 10 C is anomalous, as
it increases while increasing contact pressure. Nevertheless, some interesting

91

FRICTIONAL BEHAVIOR OF STEELPTFE INTERFACE


(%)

(a) 4

log@28.1MPa
log@9.36MPa

log@18.72MPa
P= 9.36 MPa
25
20

15
10

P= 18.72 MPa
5

0.000

5 0.000

1 00.000 1 5 0.000

2 00.000 2 5 0.000

3 00.000 3 5 0.000
(m
m
ec)
/s

P= 28.1 MPa

(mm/sec)
0
0

100

150

200

(%)

(b)

50

250

300

350

P= 9.36 MPa
25
20
15
10

P= 18.72 MPa
5

0.000

log@28.1MPa

5 0.000

1 00.000 1 5 0.000

2 00.000 2 5 0.000

3 00.000 3 5 0.000
(m
m
ec)
/s

P= 28.1 MPa

log@9.36MPa

log@18.72MPa
(mm/sec)

0
0

100

150

200

250

300

350

(%)

(c) 4

50

P= 9.36 MPa
25
20
15
10

P= 18.72 MPa
5

0.000

5 0.000

1 00.000 1 5 0.000

2 00.000 2 5 0.000

3 00.000 3 5 0.000
(m
m
ec)
/s

P= 28.1 MPa

log@28.1MPa
2

log@9.36MPa

log@18.72MPa
(mm/sec)
0
0

50

100

150

200

250

300

350

Figure 10. Variation of the friction coefcient with sliding velocity, air temperature
and bearing pressure, for lubricated interfaces. Comparison between analytical laws
and experimental results. Air temperature equal to: (a) 10 C, (b) 20 C and
(c) 50 C.

92

M. DOLCE ET AL.

observation can be made by referring to the average values of the friction


coefcient, over the three contact pressures. The friction coefcient under
service conditions (i.e. at very low sliding velocities) at 20 C air temperature is of the order of 1.52%. This is consistent with the values provided
by AASHTO under the same conditions, ranging between 2 and 2.8%, for
bearing pressures from 10 to 30 MPa. The sensitivity of the friction coefcient to temperature variations resulting from the experimental data also
conrm the conservativeness of the property modication factors suggested
by AASHTO. They are max,t = 1.5 at 10 C and min,t = 1 at 50 C, while
the experimental data under consideration provide values equal to 1.27
(0.09) and 0.59 (0.07), respectively. CEN directly provides the maximum
value of the friction coefcient to be used for verication of the bearing
and the structure under service conditions. For expected extreme air temperatures lower than 5 C, the above said design values ranges between
3% (contact pressure 30 MPa) and 6% (contact pressure = 10 MPa), thus
resulting about 23 times greater than the observed experimental values.
Based on the present experimental study, therefore, the code recommendations concerning the effects of extreme temperatures on the sliding friction
coefcient of lubricated steel-PTFE interfaces appear to be enough or even
too conservative. For lubricated interfaces, the friction coefcient is much
less sensitive to sliding velocity than for non-lubricated interfaces, as the
values of the percent ratio = (max min )/min is of the order of just
25% at 10 C and 20 C, and 40% at 50 C.
3.6. Effect of cycling
Cycling produces a double effect: (i) a reduction in the sliding friction
coefcient, due to self heating of the steelPTFE interfaces and (ii) wear
of PTFE.
The rst effect has been already discussed in 3.1. It has been found
that higher sliding rates and pressures cause a larger decrease of the friction coefcient. The decay, however, follows a negative exponential trend: it
becomes smaller and smaller while increasing the number of cycles. Moreover, the reduction in the sliding friction coefcient is only temporary: it
returns to its original value when interrupting the cycling.
As said before (see Section 2.3), the PTFE pads were changed after
every three series of tests, at the same air temperature. The thicknesses
of the PTFE pads were then measured, on at areas away from dimpled
recesses. As expected, the greatest wear of PTFE was observed at 10 C,
for interfaces with no lubrication. In this case, the thickness of the four
PTFE pads reduces by about 10%, passing from 5.4 to 4.9 mm, after having covered more than 30 m at different pressures and sliding velocities. On
the contrary, the wear of lubricated PTFE was practically negligible.

FRICTIONAL BEHAVIOR OF STEELPTFE INTERFACE

93

The wear is not a problem under seismic conditions, as an earthquake


produces only a few cycles at the maximum displacement amplitudes. It
could be important, instead, under service conditions. Actually, long-term
tests are required by different codes and prescriptions, in order to appreciate the wear of PTFE due to slow movements resulting from both imposed
thermal displacements and live loads. AASHTO, for instance, requires that
at least 50% of the usable PTFE thickness must remain after completion of
the wear test, consisting in 1.6 Km at the design contact pressure, air temperature of 20 8 C and sliding velocity not less 63.5 mm/min.
4. Modelling of SteelPTFE Sliding Bearings
Two mathematical models of the frictional behavior of PTFEsteel interfaces have been implemented.
According to Constantinou et al. (1990), the coefcient of friction at
sliding velocity v, can be approximated by the following equation:
= max (max min ) ev

(2)

in which max is the coefcient of friction at high velocities, min is the


coefcient of friction at very low velocities and is constant for a given
pressure, temperature and condition of interfaces. The frictional force is
then assumed equal to:
Fr = W Z

(3)

where W is the normal load, Z is a dimensionless hysteretic quantity which


can be calculated by solving the well-known differential equation proposed
by (Wen, 1976) in random vibration studies of hysteretic systems.
The alternative proposed analytical law to describe the relationship
between sliding friction coefcient and velocity v is derived from (Chang
et al., 1990) and is given by:

= a + b ln(v) v > 3 mm/s
(4)
= a + b ln(3) v 3 mm/s
where a and b are constant for a given pressure, temperature and condition
of interfaces. The frictional force (Fr ) is then expressed by Equation (3).
All the model parameters (i.e. max , min and on one hand, a and b
on the other hand), have been analytically expressed as a function of P
(contact pressure) and T (air temperature), through a second-order polynomial model:
f (P , T ) = 1 + 2 T + 3 T 2 + 4 P + 5 P 2

(5)

94

M. DOLCE ET AL.

The coefcients i have been obtained from a multivariate nonlinear regression, using a statistical analysis package (SPSS, 1999).
The good t of the regression, for both analytical laws, is apparent
in Figure 5, which refers to non-lubricated interface. The proposed logarithmic model, however, is more accurate in capturing the experimental
behavior in the low velocity range than the Constantinous model, which
underestimates the friction coefcient in that range.
For lubricated PTFEsteel interfaces (see Figure 10), the accuracy of the
model predictions is decidedly worse, due to the great scatter in the experimental results.
Tables II and III show the values of the model parameters for the nine
combinations of bearing pressure and air temperature considered during
the tests on non-lubricated and lubricated interfaces, respectively. In the
tables there are also compared the experimental and analytical maximum
friction forces at high peak velocities (160 mm/s, precisely). As can be seen,
the error is at most equal to 5% for non-lubricated interfaces, while it
ranges between 1% and 43% for lubricated sliding interfaces. It should be
considered, however, the different order of magnitude of the frictional force
for non-lubricated and lubricated interfaces, whose average values are equal
to about 2.5 and 0.45 kN, respectively. Thus, a 5% error implies an absolute error of 0.125 kN, in the rst case, while a 43% error implies, in the
second case, an absolute error of 0.193 kN. In any case, the maximum frictional force provided by the model can be reliably used in seismic isolation
design.

Table II. Model parameters relevant to non-lubricated interfaces, for nine different combinations of bearing pressure and air temperature values.
Experimental test
P
(MPa)

T
( C)

Fra

9.36
9.36
9.36
18.7
18.7
18.7
28.1
28.1
28.1

10
20
50
10
20
50
10
20
50

Constantious model
max
(%)

(kN)

min
(%)

1.82
1.70
1.60
2.72
2.49
2.20
3.27
2.95
2.74

8.43
6.68
6.12
6.23
4.49
3.93
4.87
3.13
2.56

19.61
18.40
17.76
13.91
12.70
12.10
11,48
10.26
9.66

0.020
0.018
0.013
0.022
0.020
0.015
0.024
0.022
0.017

Logarithmic model
Fra

(kN)

Er.
(%)

Fra
(kN)

Er.b
(%)

1.92
1.78
1.64
2.74
2.47
2.27
3.40
3.02
2.76

5
5
3
0
0
3
4
2
1

0.024
0.026
0.026
0.015
0.018
0.018
0.014
0.015
0.016

0.058
0.035
0.031
0.055
0.030
0.016
0.039
0.021
0.006

1.80
1.68
1.59
2.64
2.39
2.20
3.27
2.90
2.62

1
1
1
3
4
0
0
2
4

Comparison between experimental and analytical maximum frictional force at high peak
velocities (160 mm/s precisely)
a
Peak friction force at 160 mm/s maximum sliding velocity.
b
Percent error between experimental and numerical peak friction force.

95

FRICTIONAL BEHAVIOR OF STEELPTFE INTERFACE

Table III. Model parameters relevant to lubricated steelPTFE interfaces, for nine different
combinations of bearing pressure and air temperature values.
Experimental test

Constantious model

Logarithmic model

P
(MPa)

T
( C)

Fra
(kN)

min
(%)

max
(%)

Fra
(kN)

Er.b
(%)

Fra
(kN)

Er.b
(%)

9.36
9.36
9.36
18.7
18.7
18.7
28.1
28.1
28.1

10
20
50
10
20
50
10
20
50

0.23
0.26
0.16
0.57
0.45
0.27
0.98
0.59
0.54

2.64
1.93
1.21
2.15
1.43
0.72
2.04
1.32
0.61

3.30
2.47
1.77
2.95
2.12
1.42
3.12
2.29
1.59

0.012
0.013
0.004
0.017
0.018
0.009
0.026
0.026
0.017

0.32
0.24
0.14
0.58
0.42
0.24
0.93
0.68
0.46

41
9
15
1
8
10
5
15
16

0.002
0.001
0.001
0.001
0.000
0.000
0.002
0.002
0.002

0.022
0.017
0.009
0.021
0.016
0.009
0.020
0.015
0.007

0.32
0.24
0.16
0.81
0.55
0.32
0.97
0.71
0.47

43
10
2
42
21
18
1
19
13

Comparison between experimental and analytical maximum frictional force at high peak
velocities (160 mm/s precisely)
a
Peak friction force at 160 mm/s maximum sliding velocity.
b
Percent error between experimental and numerical peak friction force.

The accuracy of the proposed model in capturing the actual frictional behaviour of steel-PTFE sliding bearings is conrmed by Figures 11 and 12, which compare the experimental and numerical force
displacement loops of non-lubricated and lubricated interfaces, respectively. The experimental force-displacement relationships shown in Figures 11 and 12 refer to the second cycle of the tests No. 13 and
16 of Table I, at low and very high peak sliding velocities (i.e. about
15 mm/s and about 316 mm/s, respectively). Both tests have been conducted with the same displacement amplitude (50 mm), contact pressure (18.72 MPa) and air temperature (20 C). In the construction of
the numerical relationships of Figures 11 and 12, reference was made
to the displacement-time histories, as drawn from the experimental output.
As far as non-lubricated interfaces are concerned, the accordance
between experimental observations and model predictions is almost perfect,
especially at low velocities. At high velocities, the numerical model is not
able to capture the decay of the friction coefcient due to self-heating, as
expected. Larger differences between experimental and numerical results are
observed for lubricated interfaces, due to the big scatter in the experimental outcomes and to the less sensitivity of the model to velocity variation
during the applied sinusoidal displacement.
As previously noted, the gap between minimum and maximum sliding
friction coefcient is signicant for non-lubricated interfaces, while being
negligible for lubricated steelPTFE interfaces. Figures 11 and 12 clearly
prove this. As a consequence, it can be said that the behavior of lubricated

96

M. DOLCE ET AL.

Figure 11. Comparison between experimental and numerical forcedisplacement loops


of non-lubricated interfaces, at low and very high sliding velocities. Experimental tests
conducted at 50 mm displacement amplitude, 18.72 MPa contact pressure and 20 C air
temperature.

Figure 12. Comparison between experimental and numerical forcedisplacement loops


of lubricated steelPTFE interfaces, at low and very high sliding velocities. Experimental tests conducted at 50 mm displacement amplitude, 18.72 MPa contact pressure and
20 C air temperature.

steelPTFE sliding bearings tends to be the same under seismic and service
conditions, while considerable differences, in terms of maximum force and
energy loss, are found for non-lubricated steelPTFE sliding bearings.

FRICTIONAL BEHAVIOR OF STEELPTFE INTERFACE

97

5. Conclusion
A comprehensive program of experimental tests on unlled PTFE pads
sliding against polished stainless steel has been carried out. The effects
on the sliding friction coefcient of (i) type of test, (ii) sliding velocity,
(iii) contact pressure, (iv) air temperature, (v) conditi on of interfaces (i.e.
lubricated or not) and (vi) number of cycles have been investigated. The
following most important experimental ndings have been obtained:
(1) The coefcient of friction increases rapidly with velocity, up to a certain velocity value, beyond which it remains almost constant. Such value
is around 150 mm/s, regardless air temperature and bearing pressure. The
maximum velocities occurring in steelPTFE sliding bearings under an
earthquake are surely greater than 150 mm/s. Therefore, the design value of
the frictional force in seismic applications can be assumed to be independent from frequency of loading and displacement amplitude.
(2) The sliding friction coefcient of steelPTFE interfaces reduces while
increasing pressure. The reduction rate, however, depends on both sliding velocity and air temperature. It increases while increasing velocity and
while decreasing air temperature. By referring to 20 C air temperature,
18.7 MPa contact pressure and sliding velocities 150 mm/s, maximum
variations in the frictional force of steelPTFE sliding bearings of the order
of 30% are expected for 50% variations of the contact pressure, regardless
the state of lubrication of the interfaces.
(3) The sliding friction coefcient decreases while increasing the air temperature. Its rate of reduction is greater when passing from low-to-medium
temperatures, than when passing from medium-to-high temperatures. Moreover, it depends on sliding velocity, while being practically independent
from contact pressure. At sliding velocities of interest for seismic applications, the reduction rate of the friction coefcient with temperature is of
the order of 0.150.3%/ C, for non-lubricated interfaces. As a consequence,
upper and lower bound analysis is needed, in order to evaluate the maximum forces on the structural elements and the maximum displacements in
the isolation system. To this end, reference can be made to the lambda-factors (max,t and min,t ) provided by AASHTO, to estimate the values of the
friction coefcient at extreme design temperatures. Based on the experimental results of the present study, the lambda-factors suggested by AASHTO
lead to very accurate predictions for non-lubricated interfaces (differences
less than 10%), while they result too conservative for lubricated interfaces
(overestimations up to 40%).
(4) The coefcient of friction tends to decrease during continuous
loading cycles at high velocities, due to self-heating of the sliding interfaces. The phenomenon is exhausted in a few cycles, due to the attainment of a new thermal equilibrium with the ambient. Based on the

98

M. DOLCE ET AL.

available experimental outcomes, the overall decay of friction with continuous loading cycles is of the order of 2530%, for non-lubricated
interfaces.
(5) Lubrication considerably reduces (by 58 times, depending on temperature) the frictional resistance of steel-PTFE sliding interfaces and, as
a consequence, the wear of PTFE. In addition, the use of silicone grease
strongly diminishes the gap between maximum and minimum friction
coefcients and, then, the differences in the structural response between
seismic and service conditions. On the other hand, the use of silicone
grease increases the sensitivity to temperature variations of the mechanical
behaviour of steelPTFE sliding bearings.
A mathematical model of the frictional behavior of steelPTFE sliding
interfaces has been presented, which takes into account the dependence of
the frictional force on sliding velocity, contact pressure and air temperature.
The proposed model describes the relationship between friction coefcient and sliding velocity through a logarithmic analytical law. The model
is characterised by two parameters, which are analytically expressed as a
function of contact pressure and air temperature, through a second-order
polynomial equation.
The accuracy of the proposed model in capturing the actual frictional
behaviour of steelPTFE sliding bearings has been veried by comparing
the experimental and numerical forcedisplacement loops of non-lubricated
and lubricated interfaces. It has been proved that the maximum frictional
force provided by the model can be reliably used in the design of seismically isolated structures.

Acknowledgements
The authors are indebted with Ing. Roberto Marnetto (TIS SpA), and
Mr Domenico Nigro (University of Bailicata) which have cooperated in the
setting up of the testing apparatus. This work has been partially funded by
MIUR, COFIN 2002.
References
AASHTO American Association of State Highway and Transportation Ofcials (1999)
Guide Specications for Seismic Isolation Design, 2nd Edition. American Association of
State Highway and Transportation Ofcials, Washington, DC.
CEN Comite Europeen de Normalisation. TC 167 Structural bearings (2000) EN 1337
Structural bearings, Part 2: Sliding elements, Brussel, Belgium.
BS British Standards Institution (1983) BS 5400: Steel, concrete and composite bridges,
London, UK.

FRICTIONAL BEHAVIOR OF STEELPTFE INTERFACE

99

Computers and Structures Inc. (2002) SAP2000 Analysis Reference Manual, Version 8.0,
Berkeley, CA.
Constantinou, M.C., Caccese, J. and Harris, H.G. (1987) Friction characteristics of PTFE
steel interfaces under dynamic conditions. Earthquake Engineering and Structural Dynamics. 15(6), 751759.
Constantinou, M., Mokha, A. and Reinhorn, A.M. (1990) PTFE bearings in base isolation:
modelling. Journal of Earthquake Engineering 116(2), 455472.
Chang J.C., Hwang J.S. and Lee G.C. (1990) Analytical model for sliding behaviour of
Teon-stainless steel interfaces. Journal of Engineering Mechanics 116, 27492763.
Deb, S.K. and Paul, D.K. (2000) Seismic response of buildings isolated by sliding-elastomer
bearings subjected to bi-directional motion. Proceedings of the 12th World Conference on
Earthquake Engineering, Auckland, New Zealand.
Dolce, M., Cardone, D. and Marnetto, R. (2000) Implementation and Testing of Passive
Control Devices Based on Shape Memory Alloys. Earthquake Engineering and Structural
Dynamics. (29), 945958.
Dolce, M. (2001) Remarkable design examples concerning recent applications of innovative
anti-seismic techniques to bridges and viaducts in Europe. Proceedings of the 7th International Seminar on Seismic Isolation, Passive Energy Dissipation and Active Control of
Structures, Assisi, Italy.
Fan, F.G., Ahmadi, G. and Tadjbakhsh, I.G. (1988) Base isolation of a multistory building under harmonic ground motion A comparison of performances of various systems.
Tech. Report NCEER-88-0010, National Center for Earthquake Engineering, State University of New York, Buffalo.
Hwang, J.S., Chang, K.C. and Lee, G.C. (1990) Quasi-static and dynamic characteristics of
PTFE-stainless interfaces. Journal of Structural Engineering 116(10), 27472762.
Mokha, A., Constantinou, M. and Reinhorn, A.M. (1990) PTFE bearings in base isolation:
testing. Journal of Earthquake Engineering 116(2), 438454.
Mokha, A., Constantinou, M., and Reinhorn, A.M. (1993) Verication of friction model of
PTFE bearings under triaxial load. Journal of Structural Division, ASCE, 119(1), 240261.
Mostaghel, N. and Tanbakuchi, J.T. (1983) Response of Sliding Structures to Earthquake
Support Motion. Earthquake Engineering and Structural Dynamics. 11, 729748.
Mostaghel, N. (1984) Resilient-Friction Base Isolator. Report No. UTEC 84/97, Department
of Civil Engineering, University of Utah, Salt Like City, USA.
Ordinanza del PCM No 3274/2003 (2003) Primi elementi in materia di criteri generali per la
classicazione sismica del territorio nazionale e di normative tecniche per le costruzioni in
zona sismica, Roma, Italy.
SPSS Inc. (1999), SPSS Advanced models Version 9.0, Chicago, Illinois.
Takahashi Y., Iemura H., Yanagawa S. and Hibi M. (2004) Shaking table test for frictional
isolated bridges and tribological numerical model of frictional isolator. Proceedings of the
13th World Conference on Earthquake Engineering, Vancouver, Canada.
Tsopelas, P., Constantinou, M.C., Okamoto, S., Fuji, S. and Ozaki, D. (1996) Experimental
study of bridge seismic sliding isolation system. Engineering Structures 18(4), 301310.
Tyler, R.G. (1977) Dynamic tests on PTFE sliding layers under earthquake conditions. Bulletin of the New Zealand National Society for Earthquake Engineering 10(3), 129138.
Wen, Y.K. (1976) Method for Random Vibration of Hysteretic Systems. Journal of the Engineering Mechanic Division, ASCE, 102 (EM2).
Zayas, V., Low, S. and Mahin, S. (1987) The FPS earthquake protection system: experimental report. Report No. UCB/EERC-87/01, Earthquake Engineering Research Center, University of California, Berkeley.

Das könnte Ihnen auch gefallen