You are on page 1of 95

KTH ROYAL SCHOOL OF TECHNLOGY

SCHOOL OF ARCHITECTURE AND THE BUILT ENVIRONMENT

Dynamic behavior of bridge bearings


Numerical modelling of frictional behaviour of POT Bearings

GABRIEL FERNÁNDEZ BOSCHMONAR


PABLO BARDESI ISUSI

DEGREE PROJECT IN CIVIL ENGINEERING AND URBAN MANAGEMENT, 30 CREDITS


STOCKHOLM, SWEDEN 2016
i
Preface

This master thesis was performed by the authors within a collaborative agreement between
the Department of Civil and Architectural Engineering at KTH and Tyréns AB. The authors
would like to give their most honest gratitude to their supervisor Mahir Ülker-Kautsell, Ph.D.
at KTH and consultant engineer at Tyréns AB, for the challenging environment that he was
able to develop and for his ability to conduct and guide the efforts of the authors being always
willing to support them. In the same sense, both authors would like to show their gratitude
to KTH’s professor Raid Karoumi for inspiring them towards this specific field of science. The
authors are grateful as well to UPM professor José Marı́a Goicolea for providing the academic
background during previous years and being always willing to help during the thesis’s process.

At the same time, the authors would like to express gratitude to Tyréns AB for this opportunity
and specifically to the bridge division for their support. In addition, the authors want to thank
both Universities, UPM and KTH, for almost seven years of studying and for giving them the
opportunity of enriching their perspective and knowledge by allowing them to participate in
the double-degree master program.

Finally, the authors of this thesis want to thank their families, friends and loved ones for
supporting them through all of these years.

The last but not the least, Pablo Bardesi Isusi and Gabriel Fernández Boschmonar want to say
thank you to each other for such a nice experience together. This thesis is the outcome of the
deep and friendly collaboration between both of them.

Pablo Bardesi Isusi and Gabriel Fernández Boschmonar,

Stockholm 9th of February 2017.

ii
iii
Contents

Preface ii

Page

1 Introduction 1

1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1.1 Linear sliding device test . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.1.2 Pot bearing test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

1.2 Aim and Scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2 Theory 9

2.1 Contact Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.1.1 Phenomena in friction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.1.2 Mathematical models of friction . . . . . . . . . . . . . . . . . . . . . . . 11

2.1.3 Contact methods for finite element analysis . . . . . . . . . . . . . . . . 12

2.2 Contact Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.2.1 Analytical Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.2.2 Causes of the frictional heating . . . . . . . . . . . . . . . . . . . . . . . 19

2.2.3 Thermal effects in PTFE . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2.3 Constitutive models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2.3.1 Linearly elastic materials . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2.3.2 Hyperelastic materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.4 The Bouc-Wen model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

2.4.1 Analytical model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

iv
2.4.2 Parameters of the classical Bouc-Wen model . . . . . . . . . . . . . . . . 26

2.4.3 Limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3 Numerical modelling and verification against experimental data 29

3.1 Contact PTFE-Steel modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3.1.1 Models based on Dolce’s experiment . . . . . . . . . . . . . . . . . . . . 29

3.1.2 Model based on the multidirectional bearing test . . . . . . . . . . . . . 32

3.1.3 Model based on the unidirectional bearing test . . . . . . . . . . . . . . . 33

3.2 Choice of classical Bouc-Wen model parameters . . . . . . . . . . . . . . . . . . 33

3.3 Generalization of the Bouc-Wen model . . . . . . . . . . . . . . . . . . . . . . . 37

3.4 SDOF models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

3.4.1 Matlab . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

3.4.2 User-element in ABAQUS . . . . . . . . . . . . . . . . . . . . . . . . . . 39

3.5 ABAQUS: 3D implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3.5.1 Linear sliding device test implementation . . . . . . . . . . . . . . . . . . 42

3.5.2 Bearing test implementation . . . . . . . . . . . . . . . . . . . . . . . . . 42

4 Results 48

4.1 Linear sliding device . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

4.1.1 Non-lubricated contact . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

4.1.2 Lubricated contact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

4.1.3 ABAQUS: 3D Implementation . . . . . . . . . . . . . . . . . . . . . . . . 52

4.2 Bearing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

4.2.1 Multidirectional pot bearing . . . . . . . . . . . . . . . . . . . . . . . . . 54

5 Discussion 58

5.1 SDOF models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

5.1.1 Comparison between the classical and generalized Bouc-Wen models . . . 58

5.1.2 Sensitivity Analyses of the contact area for the sliding test implementation 62

5.2 ABAQUS: 3D implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

v
5.2.1 PTFE-steel contact implementation . . . . . . . . . . . . . . . . . . . . . 66

5.2.2 Bearing model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

5.3 Comparison between SDOF and 3D models for the sliding device test . . . . . . 67

5.4 Thermomechanical processes PTFE-steel contact . . . . . . . . . . . . . . . . . 70

5.5 Comparison between 3D and SDOF models for the laboratory test . . . . . . . . 72

5.6 Uni-directional pot bearing laboratory test . . . . . . . . . . . . . . . . . . . . . 72

5.6.1 Comparison with the multidirectional bearing . . . . . . . . . . . . . . . 72

5.6.2 Comparison with the linear sliding device test . . . . . . . . . . . . . . . 73

5.6.3 Unidirectional pot bearing friction modelling . . . . . . . . . . . . . . . . 75

5.6.4 Agreement between the unidirectional bearing model and new unidirec-
tional bearing test data . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

5.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

5.8 Further questions and considerations . . . . . . . . . . . . . . . . . . . . . . . . 79

Bibliography 80

vi
List of Figures

1.1 Deformation response factor and phase angle for a damped system excited by
harmonic force (from Anil K. Chopra [1]) . . . . . . . . . . . . . . . . . . . . . . 2

1.2 Testing apparatus. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.3 Details of PTFE dimpled recesses. . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.4 Description of the test rig used for the KTH laboratory test . . . . . . . . . . . 6

1.5 Photograph of the bearing as mounted in the test rig . . . . . . . . . . . . . . . 6

2.1 Static and Kinetic Friction Coefficient . . . . . . . . . . . . . . . . . . . . . . . . 10

2.2 Hysteresis Loop . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2.3 Point mass supported by a spring system . . . . . . . . . . . . . . . . . . . . . . 12

2.4 Energy representation of the mass spring system . . . . . . . . . . . . . . . . . . 13

2.5 Mass spring system under tangential loading . . . . . . . . . . . . . . . . . . . . 14

2.6 Displacement-Load diagram for frictional contact according to Coulomb . . . . . 15

2.7 Point mass supported by a spring and diagram for the Lagrange multiplier method 16

2.8 One dimension point mass supported by a spring and diagram for the Penalty
method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2.9 Percent of the average influence over the frictional heating . . . . . . . . . . . . 19

2.10 Sketch of Bouc-Wen model’s behavior . . . . . . . . . . . . . . . . . . . . . . . . 24

2.11 Influence over the hysteretic loop and the damping ration of a parameter . . . . 25

2.12 Regions of the Hysteresis Loop . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2.13 force-displacement relationship. D approximation . . . . . . . . . . . . . . . . . 27

3.1 Curves of the kinematic friction coefficients of the contact PTFE-steel . . . . . . 30

3.2 Curves of the kinematic friction coefficients of the bearing’s tests . . . . . . . . . 31

vii
3.3 Sketch of bilinear model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

3.4 Experimental hysteretic loop provided by Dolce’s experiment [10] . . . . . . . . 34

3.5 Effects over the loops by varying a . . . . . . . . . . . . . . . . . . . . . . . . . 35

3.6 Detailed effects over the loops by varying a . . . . . . . . . . . . . . . . . . . . . 35

3.7 Effects over the loops by varying the relationship of β and γ . . . . . . . . . . . 36

3.8 Detailed effects over the loops by varying the relationship of β and γ . . . . . . 36

3.9 Effects over the loops by varying n . . . . . . . . . . . . . . . . . . . . . . . . . 37

3.10 Detailed effects over the loops by varying n . . . . . . . . . . . . . . . . . . . . . 37

3.11 Sketch of the Bouc-Wen user element . . . . . . . . . . . . . . . . . . . . . . . . 39

3.12 Sketch of the Supports (from [2]) . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3.13 Simplified model for Dolce's experiment . . . . . . . . . . . . . . . . . . . . . . . 43

3.14 Model with dimples in the PTFE pads for Dolce's experiment . . . . . . . . . . 43

3.15 Detail of the dimples introduced in the PTFE pad . . . . . . . . . . . . . . . . . 44

3.16 First bearing model cut in half by a vertical plane . . . . . . . . . . . . . . . . . 46

3.17 Simplified bearing model cut in half by a vertical plane . . . . . . . . . . . . . . 47

4.1 MatLab model of the modified Bouc-Wen model for a velocity of 15 mm/s with
non-lubricated contact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

4.2 Matlab model of the modified Bouc-Wen model for a velocity of 316 mm/s with
non-lubricated contact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

4.3 User element modelization of the modified Bouc-Wen model for both 15 mm/s
and 316 mm/s with non-lubricated contact . . . . . . . . . . . . . . . . . . . . . 49

4.4 Comparison between SDOF models and experimental data for 15 mm/s with
non-lubricated contact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

4.5 Comparison between SDOF models and experimental data for 316 mm/s with
non-lubricated contact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

4.6 SDOF modelization of the modified Bouc-Wen model for a velocity of 15 mm/s
with lubricated contact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

4.7 SDOF modelization of the modified Bouc-Wen model for a velocity of 316 mm/s
with lubricated contact . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

4.8 Non-lubricated model for the small velocity 15 mm/s . . . . . . . . . . . . . . . 52

4.9 Lubricated model for the small velocity 15 mm/s . . . . . . . . . . . . . . . . . 52

viii
4.10 Non-lubricated model for the high velocity 316 mm/s . . . . . . . . . . . . . . . 53

4.11 Lubricated model for the high velocity 316 mm/s . . . . . . . . . . . . . . . . . 53

4.12 Contact pressure in the contact PTFE-steel before the movement of the steel
occurs in the detailed model with dimples . . . . . . . . . . . . . . . . . . . . . 54

4.13 Comparison between experimental data and bearing’s model in ABAQUS (3D)
and SDOF model using the lubricated model defined by Dolce . . . . . . . . . . 55

4.14 Comparison between experimental data and bearing’s model in ABAQUS (3D)
using the lubricated model defined by Dolce and SDOF model for the multidi-
rectional bearing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

5.1 Effects over the loops by varying a . . . . . . . . . . . . . . . . . . . . . . . . . 59

5.2 Effects over the loops by varying the relationship of β and γ . . . . . . . . . . . 59

5.3 Detailed effects over the loops by varying the relationship of β and γ . . . . . . 59

5.4 Effects over the loops by varying n . . . . . . . . . . . . . . . . . . . . . . . . . 60

5.5 Detailed effects over the loops by varying n . . . . . . . . . . . . . . . . . . . . . 60

5.6 Comparison between the experimental data, the classical and the generalized
Bouc-Wen models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

5.7 Deformation of the dimples due to a specific external load (Pressure of 18.7 Mpa
and temperature of 20 º C) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

5.8 Relative error between experimental data and SDOF model based on Dolce’s
model of friction (v = 15 mm/s) . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

5.9 Relative error between experimental data and SDOF model based on Dolce’s
model of friction (v = 316 mm/s) . . . . . . . . . . . . . . . . . . . . . . . . . . 64

5.10 Sketch of the sliding device test . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

5.11 Comparison between the experimental data, the SDOF and the 3D models for a
sliding speed of 15 mm/s with non-lubricated surface . . . . . . . . . . . . . . . 68

5.12 Comparison between the experimental data, the SDOF and the 3D models for a
sliding speed of 15 mm/s with lubricated surface . . . . . . . . . . . . . . . . . . 69

5.13 Comparison between the experimental data, the SDOF and the 3D models for a
sliding speed of 316 mm/s with non-lubricated surface . . . . . . . . . . . . . . . 69

5.14 Comparison between the experimental data, the SDOF and the 3D models for a
sliding speed of 316 mm/s with lubricated surface . . . . . . . . . . . . . . . . . 69

5.15 Stress (Compression) – Strain relationship of the PTFE . . . . . . . . . . . . . 70

5.16 Stress (Shear) – Strain relationship of the PTFE . . . . . . . . . . . . . . . . . 71

ix
5.17 Comparison between multidirectional and unidirectional bearing test . . . . . . 74

5.18 Comparison between both sliding device lubricated and non-lubricated test and
unidirectional bearing test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

5.19 Comparison between experimental data and unidirectional bearing’s model in


ABAQUS (3D) and SDOF model of the multidirectional bearing . . . . . . . . . 76

5.20 Scatter from the Laboratory test of the bearing for 1 Mpa . . . . . . . . . . . . . 77

5.21 Scatter from the Laboratory test of the bearing for 5 Mpa . . . . . . . . . . . . . 78

5.22 Scatter from the Laboratory test of the bearing for 10 Mpa . . . . . . . . . . . . 78

5.23 Results from both experiments 5 mm and 1 mm for the different pressures . . . . 78

x
List of Tables

1.1 Recommended damping ratio by Eurocode [5] . . . . . . . . . . . . . . . . . . . 3

3.1 Parameters for the Constantinou model given by Dolce . . . . . . . . . . . . . . 29

3.2 Parameters of the Bouc-Wen model used in the implementation . . . . . . . . . 37

3.3 Material parameters in Dolce’s implementation . . . . . . . . . . . . . . . . . . 42

3.4 Material parameters in the bearing implementation . . . . . . . . . . . . . . . . 45

3.5 Rubber hyperelastic parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

3.6 Coefficients of friction in the interactions . . . . . . . . . . . . . . . . . . . . . . 46

5.1 Increase of friction (%) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

xi
Chapter 1

Introduction

1.1 Background

Structural dynamics in general is a time-dependant analysis which covers the behaviour of


structures under the action of dynamic loads. Dynamics in railway bridges is a subject in
current expansion that encompasses different disciplines such as structural, geotechnical and
mechanical engineering. The repetitive nature of the loads introduced by a the trains can
produce a resonance state in the bridge. This is obviously not a new topic, since the first
railway bridges date from the mid-nineteenth century, but the impressively fast growth in
terms of loads and speed, especially since the emergence of the high-speed railways makes it a
trending topic within the bridge engineering. This growth responds to many factors, such as
economic growth, switch of transportation modes by taking passengers from the air and road
transport, improvement of the safety and comfort of the users and social and environmental
advantages.

The resonance in train-bridge interaction is not typically a matter of structural integrity but
comfort of the passengers, ballast instability and derailment, increase cost of maintenance in
terms of tracks, vehicles and bridges. For this reason the bridge deck acceleration design is
considered as a serviceability limit state and not a ultimate limit state.

Normally several modes of vibration are involved in the train-bridge resonance, but in the
simplest case with just the first of them, the critical speed for resonance can be determined as:

υc = f1 λi (1.1)

Being f1 the first natural frequency of the unloaded structure and λi the principal wavelength
of frequency of excitation, e.g. regular axle distance. The response will depend on the bridge
type, the complexity of the structure and design stage.

In any structure and particularly in bridges, the damping has a crucial role in its response. A
very general concept of the dynamic response is well explained in text-books such as reference
[1]. The deformation response factor corresponds to the relation between the displacement in
a static state (ust )0 and a displacement in a dynamic state u0 :

1
Figure 1.1: Deformation response factor and phase angle for a damped system excited by
harmonic force (from Anil K. Chopra [1])

u0 1
Rd = =p (1.2)
(ust )0 [1 − (ω/ωn )2 ]2 + [2ζ(ω/ωn )]2

2ζ(ω/ωn )
φ = tan−1 (1.3)
a − (ω/ωn )2

being ω the frequency of excitation of the system and ωn the natural frequency of the system,
ζ the damping of the system and φ the phase angle or phase lag.

Figure 1.1 shows the importance of this deformation response factor Rd , specially when the
excitation frequency ω is close to the natural frequency ωn and what is really relevant in this
study, the importance of the damping ζ in the control of this deformation response factor Rd .

Typically, the damping ratios in current regulations are based on tests on existing bridges,
and the dissipation of energy in the structures is modelled by viscous damping. When direct
integration of the equations of motion is needed, it is common practice to use Rayleigh damping
to compute the viscous damping matrix. If this problem is treated in a local way, the Rayleigh
damping coefficients are assigned in correspondence to the different materials/elements in the
model.

2
Table 1.1: Recommended damping ratio by Eurocode [5]

ξ Lower bound of critical damping [%]


Bridge type Span length L ≤ 20 m Span length L > 20 m
Steel and composite ξ = 0.5 + 0.125(20 -L) ξ = 0.5
Pre-stressed concrete ξ = 1.0 + 0.070(20 -L) ξ = 1.5
Filler beam and reinforced concrete ξ = 1.5 + 0.070(20 -L) ξ = 1.5

The design code chooses a conservative damping ratio based on a lower bound value from
experimental tests performed by the European Rail Research Institute (ERRI) [4], see the
damping ratios recommended for dynamic analysis of railway bridges for high-speed traffic by
Eurocode [5] in the Table (1.1). Therefore, there are other sources of non-linear damping that
has not been fully determined that are not taken into account by the current regulations due to
that conservative approach. Then, by having a better knowledge of those sources, it could be
possible to improve the design codes by optimizing the determination of the damping ration,
introducing a more realistic energy dissipating mathematical or numerical models, as it was
advanced by the authour in [3]. This sources of ”extra” damping are the track superstructure
(including loss of energy in ballast and opening and closing of cracks), friction in joints, dynamic
soil-structure interaction (SSI, see [6, 7, 8]) and finally the motivational topic of this thesis, the
friction in the bearings. According to [3, 9], the most likely sources of the mentioned non-linear
behaviour are the longitudinal track resistance and the pre-rolling resistance of roller bearings.
This is what have motivated the study of the dissipation of energy at the bearings.

Of course, from a safety point of view is important not to overestimate this damping, but
at the same time not to underestimate since it will increase the cost of construction on new
bridges and also the cost of maintenance on existing bridges. As it is well-known within the
maintenance practice community regarding maintenance operations: “the less the better”. Any
time that a maintenance operation is performed it exists either the possibility of damaging or
affecting in unexpected ways the overall behaviour or the structure.

1.1.1 Linear sliding device test

One of the main foundations of this thesis is a previous study performed by Dolce et al. [10].
Within that study, the contact between the Polytetrafluoroethylene (PTFE) or more commonly,
Teflon, and steel was tested. The variables that were considered to determine the behavior of
the contact PTFE-steel were: contact pressure, temperature, sliding velocity and the addition
or not of lubricant at the contact.

The testing apparatus consisted of a central steel plate which was sandwiched by four PTFE
pads. The central steel plate and the PTFE pads were clamped by two steel plates compressed
by a 50 KN hydraulic jack and four steel-rods which were pre-stressed. The sliding velocity was
introduced by a 10 KN dynamic actuator driving the central steel plate. In addition, the air
temperature was controlled by a K-type thermocouple, which was located at the thermal room
where the experiment was carried out.

Two main types of experiment regarding this topic could have been performed: displacement
control and load control. The first one enforces a non-zero displacement constraint and the
second one does the same but with the load instead. The reason behind using the displacement

3
control type is that the aim of this experiment was determining the kinematic friction coefficient
of the PTFE against steel and in order to do that the velocity field, i.e. the derivative of the
enforced displacement field, is required. This reason motivated the use a displacement control
test. Figure ?? shows a sketch of the testing apparatus.

The PTFE pads are not flat surfaces. The reason is that they are generally obtained from a
sheet with dimpled recesses. The dimensions can be seen in Figure 1.3. The purpose of the
dimples are retaining the silicone grease at repose and gradually releasing the lubricant over
the contact surface while sliding.

Different tests were performed varying the values of the considered variables. In the case of the
contact pressure, three different quantities were examined: 9.36 Mpa, 18.72 Mpa and 28.1 Mpa.
On the other hand, the temperature was analyzed for other three different magnitudes: -10, 20
and 50 ºC. The range for the sliding speed was from 15 mm to 316 mm. Finally, the displacement
were 10, 25 and 50 mm and both triangular and sinusoidal curves were used for the enforced
displacement.

1.1.2 Pot bearing test

Not only previous papers written by M.Ülker-Kaustell were the inspiration and background
for this work but also a current laboratory test carried out in the Structural Department of
the Royal Institute of Technology (KTH). He and his team are developing a simple model of
a pot bearing using the smallest commercial one and assuming negligible scale effects. It is
considered a preliminary test so some limitations must be accepted, the uni-directional bearing
is unstable in the direction of the bearing motion because there is no lateral support for the
mechanisms that apply the vertical load except for the friction force in the bearing. Therefore
an horizontal cylinder was used to stabilize the system and apply the motion, turning it in a
displacement-control test. Figures 1.4 and 1.5 show the mechanisms and he actual test.

In addition, it is important to consider the eccentricity between the elastic line of the structure
and the rotation center of the bearing since it influences the relation between the horizontal
displacement and the rotation over the bearing. This influence was studied by [9] and it increases
for bigger eccentricities, therefore the experiment was implemented with an eccentricity of 1
meter.

The vertical hydraulic cylinder produces forces from 10-200 KN which corresponds to contact
pressures from 1-20 Mpa approximately in the PTFE surface. The motion will vary in the order
or millimeters and with frequencies of 0.01-5 Hz.

Finally, in a normative sense the test follows the Swedish Standards (SIS) which are directly the
European Standards. This norms will be basically the Swedish Standards for Sliding elements
[11] and for Pot bearings [12]. The first of them will specify general issues for sliding bearings
such as the general conditions for the friction or the lubricants while the second of them will
provide particular conditions for the pot bearing test itself.

4
HYDRAULIC PRESTRESSING
JACK RODS

PTFE LOAD
CELL

STEEL

THERMAL
CHAMBER
LOAD CELL
DYNAMIC
ACTUATOR

Figure 1.2: Testing apparatus.

15

R4
25
13

40

8
5,4

1,3

Figure 1.3: Details of PTFE dimpled recesses.

5
Figure 1.4: Description of the test rig used for the KTH laboratory test

Figure 1.5: Photograph of the bearing as mounted in the test rig

6
1.2 Aim and Scope

The main aim of this thesis is to generalize the Bouc-Wen model of hysteresis (see Section
2.4) ) regarding its implementation in the dynamic analysis of bridges, in particular its effect
over the whole behaviour of the bearing and its interaction with the bridge through analysing
and understanding the modelling details. By doing so, this thesis also aspires to be a guiding
point for further research concerning the identification of one of the sources of the non-linear
dynamics of bridges: the hysteresis effect at bearings.

This document will explain and analyse the different contact interactions as well as the overall
mechanical behavior that exist at pot bearings, specially focusing on the PTFE-steel contact
where the hysteresis takes place. As a result of that, different models will be developed that
simulate those mechanical processes from a general perspective.

The commercial FE-code ABAQUS [13] widely used in many fields of engineering, will be used
to perform 3D simulations of the bearing test as well as the linear sliding device test. The
purpose of this process is to further increase the understanding of the bearing mechanism and
the phenomena related to it.

At the same time, single degree of freedom (SDOF) models will be used with a generalized
coordinate using Matlab and an ABAQUS 's user element subroutine in order to simulate the
hysteretic effects produced both by the bearing test and the linear sliding test. The generalized
coordinate represents the fundamental mode of vibration of the bridge in the case of the bearing
and the sliding direction in the case of the linear sliding device. The results obtained with
these simplified models made in Matlab will be compared with the ones produced by the more
complex ones defined in ABAQUS as well as with the experimental data mentioned above.
This comparison will lead to a better understanding of the dissipative effects that take place
at bearings.

7
8
Chapter 2

Theory

2.1 Contact Mechanics

The contact problem has vital importance in this research. The different materials involved in
the experimental tests considered within this thesis give the possibility of using many different
models and subsequently increases the difficulty of a proper selection. Of course this is not a
new issue, many authors have written throughout history, since the first models developed by
well-known scientists such as Amontons and Da Vinci, Euler [15] who introduced the common
used symbol for the friction coefficient µ or the well-known Coulomb's friction law [14]. The
three laws described by Amontons are.

1. The force of friction is directly proportional to the applied load.

2. The force of friction is independent of the apparent area of contact.

3. Kinetic friction is independent of the sliding velocity (Coulomb 'law)

In addition the friction force can be described in its simpler form as:

Ft = µ · N (2.1)

The development of these theories has grown hugely, especially since the appearance of the
finite element method combined with the power of the modern computers. In order to study
the frictional contact between the different surfaces of the bearing the next section will introduce
the solution methods considered in the solution of this particular problem in ABAQUS.

More particularly the use of the contact PTFE-Steel has gained ground in the last decades in
bridge bearings applications. They have been used during the past forty years to accommodate
effects such as thermal movements, pre-stressing, creep and shrinkage while in the latest years
their application has grown in the seismic isolation systems. There is an extensive literature
in the study of those bearings, analysis of the seismic response of bearings equipped with
steel-PTFE sliding isolation devices ([16], [17]), but they used the Coulomb friction law, which
it is known to be inaccurate now. Several studies of the nineties showed the dependency
on the sliding velocity and contact pressure of the friction coefficient [18], [19], [20]. The

9
temperature was revealed as an important factor too [21]. Based in the viscoplasticity theory
[22] and [24] developed a mathematical method to obtain the dynamic friction coefficient with
a sliding velocity dependency. With great relevancy for this thesis, Dolce et al. [10] developed
a laboratory device to study experimentally how the friction coefficient was affected by the
variation of the sliding velocity at different contact pressures and temperatures. Finally it is
important to highlight the contribution of Lomiento et al. [29] who performed a study of a
spherical bearing capturing the stick-slip effect and cyclic effect which will be discussed in the
next section.

2.1.1 Phenomena in friction

There are several phenomena that affect in the interaction of two materials both in a micro
and a macro scale point of view. Their importance and impact on the friction coefficient vary
depending on the materials in contact.

Firstly, looking deeper in the micro-scale contact where the micro-slip is founded. This phe-
nomenon as described by [27] is produced when the movement starts between two surfaces in
contact, where locally some parts of the surfaces get stuck while others actually slide.

Then, some effects that occur in a macro-scale level must be taken into consideration. It is
well-known the existence of a static coefficient of friction, also known as the breakaway friction
coefficient (us ) and a kinetic friction coefficient (uk ). The differentiation is necessary since the
first one will govern problems with static loads or barely not movement and the second one will
lead to dynamic problems like the one that leads this thesis. The static friction is known as the
first resistance suffered while moving two surfaces in contact while the kinetic friction appears
when the displacement occurs, Figure 2.1 shows them. Normally the static friction coefficient
is greater than the kinetic friction coefficient. This phenomenon is well described by [25] and
by [26]. This break-away effect can be seen in Figure 2.2 as part of the hysteresis loop.

Static and Kinetic Friction


Coefficients

Friction Kinetic
Static
resistance region
region
(N)

Applied
Force
(N)

Figure 2.1: Static and Kinetic Friction Coefficient

10
Force [KN]

Break-away

Cycles
Stick slip

Displacement [mm]

Figure 2.2: Hysteresis Loop


Break-away

Another phenomena observed during contact is typically called stick-slip effect, see [28] but
other authors such as [29] have referred it as a break-away effect too. This behaviour is often
observed when two surfaces in contact stick (the friction between them increases) and then slip
(the friction decreases). In the particular case of a sinusoidal force this is observed where the
displacement changes direction. An example of the stick-slip effect is shown in Figure 2.2.

In addition, the temperature in the contact varies when continuous loading cycles occur and
this has an effect in the friction coefficient. This effect can be appreciated in Figure 2.2 and it
was appreciated by Loamiento et al. and Dolce et al. in their experiments.

2.1.2 Mathematical models of friction

Dolce et al. fitted two different mathematical models. The first one is defined by Constantinou
et al. [24] as an exponential model. The coefficient of friction at certain sliding velocity can be
define as:

µ = µmax − (µmax − µmin ) · e−α·υ (2.2)

µmax is the coefficient of friction when high sliding velocities are reached and µmin the coefficient
of friction at low sliding velocities (do not mistake it with the static coefficient of friction) and
α a constant of the model for an established pressure and temperature.

The second model is based on the logathimic function. This model proposed by Chang et al.
[23] has the form:


µ = a + b · ln(υ) υ > 3 mm/s
(2.3)
µ = a + b · ln(3) υ ≤ 3 mm/s

where a and b are model parameters for a determined temperature and pressure.

The modelling of this phenomenon requires generalizing the friction rule for different pressures
and velocities. Within this thesis the temperature has been set as a constant with T = 20 º
C however this variable can be introduced at any time. According to the exponential equation

11
k

u h

Figure 2.3: Point mass supported by a spring system

of the kinematic friction coefficient Equation 2.2 and 2.3, a second degree polynomial has been
developed to model the pressure dependency of the model (µmax , µmin and α in the first case
and a and b for the second model.

f(p) = a1 · p2 + a2 · p + a3 (2.4)

Note that the model parameters in both models are constant for a certain pressure (and tem-
perature).

Finally, the frictional force is given by:

Fr = µ · W · Z (2.5)

in which W is the normal load and the Z is a hysteretic variable defined by Wen [22].

2.1.3 Contact methods for finite element analysis

This section will summarize the different methods taken into consideration to solve the contact
problem, which have been masterly explained by Wriggers [30]. Even if some of them have
not been used to extract final results, they have been crucial in order to reach convergence.
One-dimensional examples will be used to illustrate the methods.

General formulation

Figure 2.3 shows a first general problem with a point mass m under gravitational load supported
by a spring with stiffness k. A rigid plane restricts the deflection of the point mass m.

The definition of the energy of this simple system will be useful for explaining the methods
available to solve the contact problem.

12
π

π min
C

π min

u
Figure 2.4: Energy representation of the mass spring system

The energy of the previous point mass system can be written as in the next equation. In
addition Figure 2.4 represents graphically the energy in the system.

1 2
Π (u) = ku − mgu (2.6)
2

If there is no restriction on the displacement u, the variation of the energy will be:

1 2
δΠ (u) = ku − mgu = 0 (2.7)
2

Which minimum as the reader can see in Figure 2.4 is found at u = mg/k.

In addition the motion of the mass is restricted by the rigid support (no penetration) by the
next inequation:

c(u) = h − u ≥ 0 (2.8)

With c(u) representing the gap between the mass and the rigid support. The variation δu is
restricted by the rigid plane Figure 2.3 therefore δu ≤ 0, meaning that the virtual displacement
can only point in the upward direction. Introducing this condition in the energy variation from
Equation 2.7 the variational inequality is obtained:

k u δu − m g δu ≥ 0 (2.9)

So if the condition from Equation 2.8 is introduced, the minimum point is not anymore Π min
and the new point with the minimum energy within the admissible solution space is Π cmin .

Contact with friction

Specifically Figure 2.5 shows the contact problem where a point mass m under gravitational
load has already make contact with the rigid plane.

13
k
kh

mg
FT RT FT
m

RN
Figure 2.5: Mass spring system under tangential loading

The contact between the mass and the rigid support leads to a normal reaction force which as
in classical contact mechanics is considered negative, therefore the contact pressure can be only
compression.

RN ≤ 0 (2.10)

In order to consider the tangential behaviour of the problem, an additional tangential force is
applied in the system. The equilibrium equations in the vertical and tangential directions are:

RN + mg − kh = 0 (2.11)

RT − FT = 0 (2.12)

The constitutive Coulomb's law provides an inequality in which two states can occur. These
two states are known as stick and sliding. In the first one there is no relative tangential
movement between the mass and the rigid support. During the second one there will be a
relative displacement between the mass and the rigid support. These states will lead to the
next equations.

1. Constitutive Coulomb's law.

f (RN , RT ) = |RT | + µRN ≤ 0 (2.13)

µ is the friction coefficient and the absolute value of the tangential force is considered
since it can have positive or negative direction as it can be seen in Figure 2.6.

14
FT

uT

Figure 2.6: Displacement-Load diagram for frictional contact according to Coulomb

2. Stick happens when


|RT | < −µRN (2.14)
In this state there is no relative tangential displacement between the support and the
mass, therefore uT = 0. In this case the tangential force RT can be calculated by Eq2.12.

3. Slip happens when


|RT | = −µRN (2.15)
In this state there is relative tangential displacement between the support and the mass
therefore uT 6= 0. The tangential force RT will have opposite direction to the direction of
the relative displacemente uT and it can be calculated with the Equation 2.15.

Now, by introducing the form of the Kuhn-Tucker complementary conditions (Eq. 2.16) in
the theory of optimization, better known as Hertz-Signorini-Moreau [48] conditions in contact
mechanics, the inequalities can be reformulated in the Kuhn-Tucker form (Eq. 2.17).

c(u) = 0, RN ≤ 0 and RN c(u) = 0 (2.16)

|uT | ≥ 0, f ≤ 0 and |uT |f = 0 (2.17)

The absolute value of the tangential force is introduced since it can have positive or negative
direction. This is also shown in Figure 2.6.

Lagrange multiplier method

The contact problem established in Equation 2.8 can be solved by the Lagrange multiplier
method. The Lagrange multiplier method consists in the addition of a term known as the
Lagrange multiplier (λ) shown in Figure 2.7.

15
k

u h

Figure 2.7: Point mass supported by a spring and diagram for the Lagrange multiplier method

1
δΠ(u, λ) = ku2 − mgu + λc(u) (2.18)
2

Comparing with the Hertz-Signori-Moreau Equation 2.16 condition, the Lagrange multiplier is
equivalent to the reaction force. The variation of energy leads to two equations since δu and
δλ can be derived independently.

k u δu − m g δu − λ δu = 0 (2.19)

c(u) δλ = 0 (2.20)

The first equation represents the equilibrium for the point mass including the reaction force
in the contact with the surface as Figure 2.7 shows and the second equation represents the
kinematic constraint equation (Equation 2.8) in the contact: u = h . In the Lagrange multiplier
method the variation is no longer restricted, therefore if Equation 2.10 is observed:

λ = kh − mg = RN (2.21)

If the contact does not remain, meaning that the condition Equation 2.10 is no longer met, the
Lagrange multiplier will be zero and the correct solution will be the minimum u = mgk
.

Penalty method

Figure 2.8 shows how a different term has been applied from the Lagrange multiplier method.
Instead of just applying a multiplier, a spring stiffness is introduced. This leads to:

16
k

u h

Figure 2.8: One dimension point mass supported by a spring and diagram for the Penalty
method

1 1
δΠ(u, λ) = ku2 − mgu + [c(u)]2 with  > 0 (2.22)
2 2

In this case the variation is unique:

kuδu − mgδu − c(u)δu = 0 (2.23)

where the next solution is derived

mg + h
u = (2.24)
k +

Introducing it in the constraint described in Equation 2.8.

kh − mg
c(u) = h − u = (2.25)
k +

The contact occurs if mg ≥ kh, therefore an equivalent compression of the spring would happen
when a hypothetical penetration of the mass would occur in the support. The constraint is
fulfilled in the limit  → ∞ ⇒ c(u) → 0. Two cases are derived:

1.  → ∞ ⇒ u − h → 0, meaning that for very large parameters the solution is correctly


approached. If the stiffness of the spring is very high, small penetration occurs.

2.  → 0 is only valid for inactive constraints, because it means that it is unconstrained. If


the stiffness of the spring is very low, high penetration occurs.

In the same way as in the Lagrange multiplier method, the reaction force is computed as
RN = c(u). Therefore using Equation 2.23:

17

RN = λ = c(u) = (kh − mg), (2.26)
k +

which in the limit  → ∞ converges in the correct solution obtained with the Lagrange
multiplier method, see Equation 2.21.

2.2 Contact Temperature

It is known that the friction processes are not satisfactory understood. Among those processes,
the thermal behavior of bodies that are sliding is considered a multifaceted problem.

Sliding bodies generate heat, which is called frictional heating; if there is an effective dissipation
of heat it generally leads to a decrease of the temperature produced by the contact. However,
in the case of polymers, frictional heating might produce that some regions at the contact
surface reach the melting or softening temperature and consequently polymer materials can
suffer a change of friction properties and wear rates, see [31] and [32]. This rise of temperature
due to the frictional interaction may also change the surface geometry, causing local wear for
instance. Within the damaged region, the capacity of dissipating heat gets decreased; this
provokes an increase in the temperature reducing the cohesive forces [33]. This effect favors the
appearance of cracks. There are several studies that state the tendency of a reduction of the
friction coefficient at the surface contact between polymers and metals while the applied load
and the sliding velocity get increased, see [34]. A study developed in [35] describes the influence
of the frictional heating over rubber materials for different values of pressures and velocities.
The study also highlights the importance of locally studying the variation of temperature. The
reason is that the viscoelastic materials have a tremendous dependence of the temperature and
the distribution of the temperature is not necessarily homogeneous.

2.2.1 Analytical Models

The temperature produced at the contact between different bodies it is generally defined as:

TContact = TBulk + TFlash (2.27)

TBulk is the bulk temperature of the body or bodies before entering in contact, which is the
air temperature. TF lash is the flash temperature, which is also called frictional heat and is
achieved almost instantly at the asperities of the contact when sliding. Several models are used
to calculate the flash temperature, however, they can be gathered in two main groups. One
uses the geometry of the contact surface while the other focuses on the material properties not
accounting the specific geometry.

The model developed in [37] represents the second type of models. The equation governing this
model is given by:

18
15 %

Roughness

35 %
Contact
pressure

Sliding
velocity

50 %

Figure 2.9: Percent of the average influence over the frictional heating

1
g µ υ (W Pm ) 2
Tflash = 0.443 (2.28)
J (k1 + k2 )

In this equation, g is the constant of acceleration (9.81 m/s2), µ is the kinetic friction coefficient,
υ is the slip-rate or the sliding speed (m/s), W is the contact load (N), P m is the hardness
of the softer material placed in contact (Mpa), J is the heat’s mechanical equivalent, which is
4.186 J. Remember that the mechanical equivalent of heat relies on the idea that motion and
heat are mutually interchangeable since any amount of work could generate the same quantity
of heat if the work done is fully converted to heat energy. Finally, k1 and k2 are the thermal
conductivities of the contacting materials (W/mK).

By using this model, the work produced by the friction force can be directly computed con-
sidering just the thermomechanical properties, remarking the strong influence of the velocity
and friction, since both appear explicitly in the formula. The other kinds of models explicitly
require the specific geometrical dimensions of the contact.

2.2.2 Causes of the frictional heating

According to [36], the main factors of the frictional heating are: the sliding velocity, the applied
pressure and the surface roughness. Additionally, the influence of each one was measured
producing the results that can be seen in Figure 2.9.

However, the study showed that the sliding velocity was the most significant factor. It was ob-
served in the study that when the sliding velocity was increased by a factor two the temperature
was almost doubled, suggesting a proportional relationship between them. The relationship
seemed to be proportional as well in the case of the contact pressure, when this factor was
doubled the contact temperature was increase by a factor of one-third. On the contrary, the
roughness did not seem to have a proportional relationship.

19
2.2.3 Thermal effects in PTFE

It is possible to state that the structure and, as a result of that, the behavior of the PTFE
varies when its temperature gets increased, see [38]. The heat provokes the appearance of
volatile constituents that increase the thermal stability and the crystallinity of PTFE. There
are different degrees for these effects on the polymer chains: they can get thinner, cut, removed,
cracked or even pulverized. The heating produced by friction causes a considerable amount of
heat in the bulk material. Thermal effects are a very significant factor to take into consideration
regarding the durability of the polymer chains, see [36].

Transfer films

The sliding contact between PTFE and steel can reach temperatures that melt parts of the
polymer. The adhesion and joining of those melted fragments that get stuck within the as-
perities of the steel creates what it is called transfer films, see [39]. These films appear at the
contact surface alleviating the contact by absorbing heat. This behavior highly influences the
mitigation of the thermomechanical properties by helping the PTFE to dissipate heat at the
contact and softening its roughness at the contact. For this reason, the contact is always more
severe at the beginning when the PTFE did not have enough time to produce this transfer
layer. This could be the reason why there is a reduction of the friction force in the hysteresis
loops due to the increase of the number of cycles, the layer takes some time to get formed and
then it softens the contact.

However, transfer films can be removed from the contact surface due to the friction forces,
leading to a significant increase of the temperature. When this layer is removed the coefficient
of friction increases until the new transfer film appears. In other words, the transfer film has a
dual effect over the contact properties of the PTFE and steel: accumulate heat and reducing
the coefficient of friction.

2.3 Constitutive models

Due to the relevancy of the material behaviour within the thesis, some constitutive models shall
be introduced.

2.3.1 Linearly elastic materials

The elastic materials are govern by the classic Hook 's law of elasticity. It can be derived from
a quadratic form of the state potential, with the Young 's modulus E and the Poisson 's ratio
ν as characteristic materials

1
Ψ∗ = Aijkl σij σkl (2.29)

∂Ψ∗ 1+ν ν
εij = ρ = σij − σkk δij (2.30)
∂σij E E

20
In addition, the so-called perfectly elastic material is a material that locally produces no entropy.
This is applied to a material when in every admissible process (excluding natural, damage,
viscous mechanisms and plastic deformations) the internal dissipation Dint is equal to zero.
Mathematically this is governed by the next equation:

∂Ψ(F)
Dint = P : Ḟ − Ψ̇ = (P − ) : Ḟ = 0 (2.31)
∂F

F is the deformation gradient, Ψ the strain-energy function and P the first Piola-Kirchhoff
tensor also known as the thermodynamic force work conjugate to F. This procedure is well
described by [41].

2.3.2 Hyperelastic materials

Some materials cannot be described as elastic materials, for instance the rubber is a elastomer
and it is considered to be an isotropic hyperelastic material see e.g. the textbook by Holzapfel
[42]). These materials are characterized by a similar response, based in a stress-strain con-
dition, regardless the direction studied which intuitively responses to the idea of a rubber.
Mathematically this corresponds to the idea that between a reference configuration (Ω0 ) and
a new configuration(Ω∗0 ) which are rotated by a´n orthogonal tensor Q the strain energy is
constant:

Ψ(F) = Ψ(F∗ ) = Ψ(FQT ) (2.32)

Therefore if the strain-energy function is equal the material is called isotropic and otherwise
it is called anisotropic. For further discussion see [43]. The material analysis of the rubber
goes further by studying the compressibility. This kind of material experiments no noticeable
change of volume even under big strains so they can be regarded as incompressible isotropic
hyperelastic materials, which introduces the following constraint:

J = detF = 1 (2.33)

J is the volume ratio. This is a material subjected to an internal constraint or so-called con-
strainted material. The strain-energy function is postulated as:

Ψ = Ψ(F) − p(J − 1) (2.34)

The scalar p corresponds to the hydrostatic pressure and it works as a Lagrange multiplier (see
Section 2.1.3) that will be determined from the equilibrium equations and boundary conditions.
Therefore first Piola-Kirchhoff stress tensor will be:

∂Ψ(F)
P = −pF−T + (2.35)
∂F

21
For incompressible hyperelasticity Ḟ is not arbitrary which means that the expression in paren-
thesis in 2.31 is not zero. The incompressibility constraint defined as J˙ = F−T governs Ḟ so
Equation 2.31 can be rewritten as:

∂Ψ(F)
(P − + pF−T ) : Ḟ = 0 (2.36)
∂F

For further detail check [44] and [45]. Now if the strain-energy is expressed in terms of invari-
ants (for instance C or b,) which are the Cauchy-Green tensors and can be defined from the
deformation gradient tensor F and the rotation tensor R:

F=RU=vR with U2 = U U = C and v2 = v v = b (2.37)

It can be said that in general:

Ψ = Ψ[I1 (C), I2 (C), I3 (C)] = Ψ[I1 (b), I2 (b), I3 (b)] (2.38)

Applied to the incompressible isotropic hyperelastic material:

1 1
Ψ = Ψ[I1 (C), I2 (C))] − p(I3 − 1) = Ψ[I1 (b), I2 (b)] − p(I3 − 1) (2.39)
2 2

Deriving respect the tensor C and expressing the constitutive equations in terms of the invari-
ants I1 and I2 :

∂Ψ ∂Ψ ∂Ψ 2
σ = −pI + 2( + I1 )b − 2 b (2.40)
∂I1 ∂I2 ∂I2

which can also be written:

∂Ψ(b) ∂Ψ(b)
σ = −pI + 2 b = −pI + 2b (2.41)
∂b ∂b

Finally, expressing the stress-strain equation in terms of the three principal stretches λa and
setting the incompressibility constraint J = 1

∂Ψ
σ = −p + λa with a = 1, 2, 3 (2.42)
∂λa

J = λ1 λ2 λ3 = 1 (2.43)

22
2.4 The Bouc-Wen model

The term hysteresis is derived from υστ ρησιζ, which is an ancient Greek word that means
“deficiency” or “lagging behind”. In this particular case, “lagging behind” refers to the depen-
dence of the state of a system on its history. In other words, in order to predict how the system
is going to behave in the next step, knowledge of its behavior in the previous steps is necessary.

In the theory of plasticity, it is used with the meaning that the stress-strain or force-displacement
curves follow different paths upon loading and unloading. In addition, the confined area pro-
duced by the hysteresis loop represents the energy dissipated by the system due to its material
internal friction. Note that hysteretic systems can recover completely their deformation if the
plastic limit is not reached and the load is completely removed while dissipating energy given
that the path of both curves (loading-unloading) is different.

A model is a theoretical approximation of the reality, which simplifies the actual system while
trying to maintain accuracy. The simulation of complex non-linear dynamic interactions usually
requires a huge computational cost; however the introduction of macro elements reduce the pre-
processing tasks while allowing to characterize quite accurately the real behavior, see [52]. In
the same sense, the models used in structural dynamics usually employ macro elements since
quite complicated structural systems can be reduced to spring and dashpot elements, structural
elements or components can be resembled by those simpler elements. In general, all the damping
models used in structural dynamics are hysteretic, however, those linear models can be modified
to include also the non-linear effects.

There are quite many Hysteresis models, such as the Bi-linear or Tri-linear models, the Ramberg-
Osgood model or the Clough degrading model. Although, the most common model within the
civil engineering field is the Bouc-Wen model. The use of this model is motivated by the follow-
ing reasons: This model is able to capture, in an analytical form, a range of hysteretic loops that
reproduce quite well the behavior of a wide group of hysteretic systems. The models previously
mentioned are simpler than the Bouc-Wen model but their ability to generate a great variety of
curves is quite limited, specially if they are compared with the Bouc-Wen model. In addition,
even if the model is more complex than others that could be considered, the model is still
reasonably simple. The Bouc-Wen model is built on the combination of two macro elements: a
linear spring and the Bouc-Wen element, which can be seen as a hysteretic or non-linear spring.
In conclusion, The Bouc-Wen model of hysteresis is simple enough to be easily implemented
and generalized while being accurately enough to simulate physical phenomena.

2.4.1 Analytical model

The restoring force in the classical univariate Bouc-Wen model is given by:

F (t) = a k0 u(t) + (1 − a) k0 D z(t) (2.44)

Here u is the displacement, a is the ratio of the plastic and initial stiffness (k0 ). D is the
plastic deformation limit and z is an internal variable, responsible for the hysteretic behavior
of the model. In other words, the model is described, as it was introduced before, by two
macroelements, Figure 2.10 shows a sketch of its behavior.

23
Elastic Postyielding Spring

Fel

Fel,u
aki
u

F, u
Hysteretic Spring

Fh
Fhmax

(1-a)ki Fh, u
u

-Fhmax

Figure 2.10: Sketch of Bouc-Wen model’s behavior

The internal variable is governed by an ordinary differential equation that defines the hysteretic
behavior, the form of which can be derived by means of the endochronic theory of plasticity,
see [53] and [54]. The non-linear ODE is given by:

ż = u̇/D (1 − (β + sign (u̇ z) γ)) |z|n (2.45)

Here β, γ and n are dimensionless parameters governing the behavior of the model and the
shape of the hysteretic loops. The function sign (x) is defined as:


 1, x > 0
sign(x) = 0, x = 0 (2.46)
−1, x < 0

However, the parameters of the Bouc-Wen model are functionally redundant because it is
possible to produce the exactly same response for an external excitation by using different
combinations of them [58].

On the other hand, it is quite interesting to analyze the influence of the parameter a over the
damping ratio because it can also be seen as the representation of the relationship between
the histeretic energy and the maximum elastic energy. The restoring force has two parts, the
first one which is linear and can be described by a spring with k0 stiffness and the second part
which is governed by the hysteretic parameter and is not linear. If the dissipative energy is

24
a = 0.001 a = 0.01 a = 0.1 a = 0.5

Figure 2.11: Influence over the hysteretic loop and the damping ration of a parameter

represented by the area inside the loop, the damping ratio for a hysteretic system can be defined
with the hysteretic energy, which is given by the line integral around the loop:

Z
Eh = F du (2.47)

And the maximum elastic energy stored during the loop:

F m um
Ee = (2.48)
2

And the damping ration, see [55], would be defined by:

Eh
ξ= (2.49)
4π Ee

Therefore, for small values of a, the damping will be governed by the hysteretic energy while
when a is close to 1 the damping will be governed by the linear elastic energy, see Figure 2.11.

Additionally, the response of the Bouc-Wen model can be divided into four different regions
depending on the sign of du/dt and z as it can be seen in Figure 2.12.

4
du/dt > 0, z > 0 du/dt < 0, z > 0
3

2
Force [KN]

-1

-2

-3
du/dt > 0, z < 0 du/dt < 0, z < 0
-4
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
Displacement [mm]

Figure 2.12: Regions of the Hysteresis Loop

25
2.4.2 Parameters of the classical Bouc-Wen model

The following parameter analysis will be performed according to physical and mathematical
consistency. The concept of BIBO stability will be used for that purpose. Remember that
BIBO stability means bounded-input, bounded-output. If a system is BIBO stable, then the
output will be bounded for every input to the system that is bounded.

k0

k0 is the elastic stiffness and can be easily determined from the experimental data.

The parameter a is defined as the ratio between the elastic and post elastic stiffness. a can not
be below 0 because for negative values the loops can not describe adequately real structural
systems [56]. The limit value of a = 0 would provoke that the response of the system would be
governed only by the hysteretic element. In addition, a > 0 is required in order to be BIBO
stable [56].

γ and β

Wen demonstrated that these two non-dimensional parameters govern the shape and size of
the loop [58]. However, there is no physical explanation behind them and the effects over the
whole response are not well measured since they influence the model in an indirect way [59].

There is only one combination of parameters β and γ, in combination with a, that will led
to BIBO stability, being at the same time compatible with the free motion of real structural
systems and the laws of thermodynamics [56] :


a>0
(2.50)
β+γ >0

At the same time, thermodynamic admissibility issues impose the subsequent inequality [57]:

γ≥β (2.51)

In addition, there is a criteria that it is usually introduced in order to reduce the Bouc-Wen
model to a strain-softening expression with well-defined properties [60]. This criteria lead to
the following equation:
β+γ =1 (2.52)

The reasons behind this inclusion were both physical and mathematical because it provokes
that the hysteretic value z varies from -1 to 1. When z = 1 yield has fully happened as a result
of a force acting in the positive direction (loading). When it takes the value of -1 yield has
happened as well but with the force acting in the negative direction (unloading). Obviously,

26
values within that range represent intermediate states between those limit cases. It is true that
the model loses versatility with this inclusion but its implementation seems much more clear
and rigorous avoiding or at least reducing the ambiguity of the parameters [59]. In addition,
the inclusion of this criteria removes the redundancy of the Bouc-Wen model parameters [58].

The exponential parameter n controls the form of the transition of the elastic and post elastic
part. Regarding its consistency within the model,

n ≥ 1 because if n ≤ 1 the local Lipschitz property is not fully verified by all the differential
equations, which can provoke that the solutions are not unique [56]. When the limit case of
n = 1 is analyzed in the same paper it is concluded that the theorems hold for this value. The
larger the value of n, the response behaves more and more similar to the bilinear model, which
means being less rounded [59].

D is denoted by the plastic displacement limit or yield displacement. Although, this is not
entirely accurate since the yield point, which is built on the yield displacement and yield force,
does not belong to the response of the Bouc-Wen model under monotonic loading. The yield
point needs to be determined by experimental observations in a real system.

Conversely, D is rather the yield point of the corresponding bilinear model [59], see Figure
2.13. The important implication of this is that the yield point can not be employed to find the
optimum yield parameters. Moreover, it is needed to use “special identification procedures” so
that the undetermined parameters can be estimated [61].

kf = aki
FY

ki= FY/D

D u

Figure 2.13: force-displacement relationship. D approximation

27
This parameter is used by the Bouc-Wen model in order to determine the family of backbone
curves that can be produced.

2.4.3 Limitations

As every model, the Bouc-Wen model of hysteresis has its advantages and disadvantages, but
among the latter ones there are two that are quite significant: It is not able to catch the
so-called “break-away” effect and the stick slip, see Section 2.1.1.

However, it can be argued that in terms of global damping the first phenomenon is not so
important from a quantitative point of view, even if it is quite important from a qualitative
point of view. The reason behind this statement is that the damping is measured by the
dissipation of energy that it is represented by the area inscribed by the hysteretic loop and in
this sense the extra area provoked by the inclusion of the “break away“ effect can be negligible
in general terms. In the case of the stick slip, the influence in terms of energy may be even
more negligible since the force is going up and down, meaning that it would balance the energy
by adding and discounting, i.e. they would sum out.

28
Chapter 3

Numerical modelling and verification


against experimental data

This chapter will describe the procedures to implement the different numerical models. First
the contact PTFE-Steel is evaluated in Section 3.1, the in Section 3.2 the choice of Bouc-Wen
model parameters is explained while Section 3.3 the new model is described. Later Section 3.4
explains the SDOF models and finally Section 3.5 shows the 3D implementation in ABAQUS.

3.1 Contact PTFE-Steel modelling

3.1.1 Models based on Dolce’s experiment

This section will explain how the information given by Dolce was interpreted. It is important
to expose that the values that were used to build the second order polynomials presented in
Equation 2.4 are the experimental values obtained by Dolce et al. [10], and that the friction
coefficient was not tested for pressures under 10 Mpa, so there is a gap for future research on
this topic.

Attending to the first model, the values provided by linear sliding device test for the given
temperature are expressed in the following Table 3.1 and this leads to the polynomials given
in Figure 3.1a for the non-lubricated case and in the Table 3.1 and the Figure 3.1b for the
lubricated case.
Table 3.1: Parameters for the Constantinou model given by Dolce

Non-Lubricated Lubricated
p [Mpa] µmin µmax α µmin µmax α
9.36 6.68 18.4 0.018 2.47 1.93 0.013
18.7 4.49 12.70 0.020 2.12 1.43 0.018
28.1 3.13 10.26 0.022 2.29 1.32 0.026

The obtained equations for the non-lubricated case, where the pressure must be introduced
in Mpa, are the following:

29
0.3
Mumax
Mumin
Alpha
0.25

0.2

0.15

0.1

(a) Non-lubricated contact


0.05

Friction coefficient and Alpha value [-]


0
0 5 10 15 20 25 30
Pressure [MPa]

30
0.3
Mumax
Mumin
Alpha
0.25

0.2

0.15

0.1

(b) Lubricated contact


0.05

Friction coefficient and Alpha value [-]

Figure 3.1: Curves of the kinematic friction coefficients of the contact PTFE-steel
0
10 12 14 16 18 20 22 24 26 28 30
Pressure [MPa]
30

30
Mumax

Mumax
Mumin

Mumin
Alpha

Alpha

25
25

20
20

Pressure [MPa]
Pressure [MPa]

15
15

10
10

5
5

0
0

0.25

0.15

0.05
0.3

0.2

0.1

0
0.25

0.15

0.05
0.3

0.2

0.1

Friction coefficient and Alpha value [-] Friction coefficient and Alpha value [-]

(a) Multidirectional bearing (b) Unidirectional bearing

Figure 3.2: Curves of the kinematic friction coefficients of the bearing’s tests

31
µmax (p) = 0.000188 · p2 − 0.011381 · p + 0.27418 (3.1)
µmin (p) = 0.000055 · p2 − 0.004010 · p + 0.10074 (3.2)
α (p) = 0.0002 · p + 0.016 (3.3)

It can be seen that both µmax and µmin tend to decrease with higher values of the pressure.
Therefore, the kinematic friction coefficient will get decrease in the same sense.

The obtained equations for the lubricated case, where the pressure must be introduced in Mpa,
are the following:

µmax (p) = 0.0000296 · p2 − 0.0012066 · p + 0.0333967 (3.4)


µmin (p) = 0.0000259 · p2 − 0.0013301 · p + 0.0301091 (3.5)
α (p) = 0.0000168 · p2 + 0.0000626 · p + 0.0109382 (3.6)

In this case µmin tends to decreases with higher values of the pressure while µmax at some point
increases again.

3.1.2 Model based on the multidirectional bearing test

Firstly, the bearing tests were modeled using Dolce’s model for PTFE-steel lubricated surfaces.
Due to the limitations that were found between Dolce’s model for the lubricated interaction
between PTFE-steel and the results obtained in the lab, it was decided to develop an adjustment
of the previous model for the SDOF multidirectional bearing models. The reason of this
limitation is that the friction coefficient can not decrease beneath the bottom limit of the
coefficient (µmin ), see Equation 2.2, which in the case of Dolce’s model is not small enough
because the sliding device test did not study slip-rates as low as the bearing test did. The
adjustment of the model is based on the multidirectional bearing test, that was performed for
2 mm of displacement’s amplitude.

Firstly, it was decided to keep using Constantinou’s equation and use the values from the
bearing test to determine the new values of µmax , µmin and α. Statistically, the best fit for the
first two parameters was obtained with a logarithmic function instead of a polynomial one, the
value of R-square of the logarithmic case was greater than 0.96 both for µmax and µmin . Then,
a second order polynomial was used to model α and Matlab was employed to determine the
best agreement with the data from a statistical point of view. The following equations show
the model functions that were obtained:

µmax (p) = − 0.02 ln (p) + 0.0777 (3.7)


µmin (p) = − 0.003 ln (p) + 0.0103 (3.8)
α (p) = 0.0002 · p2 − 0.0035 · p + 0.0698 (3.9)

The curves produced by the previous equation can be seen in Figure 3.2a. Note that the
pressures considered in the bearing tests go from 1 to 20 Mpa. This improvement was not

32
implemented in the 3D model in ABAQUS because the experimental data measures the friction
produced by the whole mechanism, not just the contact between PTFE-steel, it would not be
accurate to implement the friction produced by the whole bearing directly to a contact surface
in the 3D model.

3.1.3 Model based on the unidirectional bearing test

It was decided to keep using Constantinou’s equation (Equation 2.2) and use the values from
the bearing test to determine the new values of µmax , µmin and α. Once again, the best fit
for the first two parameters was obtained with a logarithmic function, the value of R-square
of the logarithmic case was greater than 0.98 both for µmax and µmin . Again, a second order
polynomial was used as well to define α and Matlab was employed again to determine the best
agreement with the data from a statistical point of view. The following equations show the
values that were obtained:

µmax (p) = − 0.03 ln (p) + 0.1224 (3.10)


µmin (p) = − 0.0061015 ln (p) + 0.0228620 (3.11)
α (p) = 0.0001 · p2 − 0.0028 · p + 0.0504 (3.12)

The curves produced by the previous equation can be seen in Figure 3.2b. This model of friction
for the unidirectional bearing has been implemented in the 3D model of ABAQUS. There are
two reason for this decision, first, there is no previous friction model of the contact between
the PTFE and the guide and second it is a way to test how sensitive is the 3D model to the
PTFE surface interaction definition.

3.2 Choice of classical Bouc-Wen model parameters

The bilinear model of hysteresis is the simplest mathematical model to simulate hysteretic
behavior. The force-displacement relationship is composed of piecewise linear and continuous
relationships. Initial or elastic loading and unloading follow straight lines with slope k0 and the
post initial or post elastic loading follows straight lines with slope a k0 , see Figure 3.3.

For certain values of γ, β and n, the Bouc-Wen model tends to reproduce loops quite similar
to the ones produced by the bilinear model. In order to simplify the choice of those Bouc-Wen
model parameters they have been selected to behave like the bilinear model.

33
F
ak0

k0 k0

k0

ak0

Figure 3.3: Sketch of bilinear model

1
Force [N]

-1

-2

-3

-4
-0.05 -0.04 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04 0.05
Displacement [mm]

Figure 3.4: Experimental hysteretic loop provided by Dolce’s experiment [10]

k0

The elastic stiffness should be relatively high, since the slope of the elastic part of the loop,
i.e. the elastic stiffness, is quite steep in the experimental curves. Moreover, the value can be
unequivocally determined in the case of the bearing test where the beginning of the loop is
provided, see Table 3.2. On the other hand, in the case of the sliding device test where the
beginning of the curve is not provided the initial can be determined by relying on the slope of
the unloading and elastic loading branches of the loops.

What can be seen from the bearing test and the linear sliding device test is that the post
elastic region of the loops is quite flat, see Figure 3.4. Figure 3.5 shows how the slope varies
for different values of a.

For values of a < 10−4 , no significant variation in the shape of the hysteresis loop is obtained.
Therefore, when a is lower than a certain value, which in the considered cases is around a =
10−4 , the loops are not so sensitive to this parameter anymore. It was decided to use a = 10−12 .

34
a = 1e-3 a = 1e-4 a = 1e-5 a = 1e-12
30

20

10
Force [KN]

-10

-20

-30
-50 -40 -30 -20 -10 0 10 20 30 40 50
Displacement [mm]

Figure 3.5: Effects over the loops by varying a

a = 1e-3 a = 1e-4 a = 1e-5 a = 1e-12


25

24.5

24
Force [KN]

23.5

23

22.5

22

-3 -2 -1 0 1 2 3
Displacement [mm]

Figure 3.6: Detailed effects over the loops by varying a

Finally, in the case of 10−3 the loops becomes sensitive to this parameter because of the long
sliding distance considered in sliding device test but for smaller sliding distance such as the
ones considered in the bearing test, between 0.01 and 5 mm, the influence is reduced quite
much, see Figure 3.6. Note that the variation between a = 10−3 and the rest of the curves is
less than 2 % for 3 mm.

γ and β

Figure 3.7 and 3.8 show how γ and β parameters affect the hysteretic loop.

35
Gamma = 0.1, Beta = 0.9 Gamma = 0.5, Beta = 0.5 Gamma = 0.9, Beta = 0.1
30

20

10
Force [KN]

-10

-20

-30
-50 -40 -30 -20 -10 0 10 20 30 40 50
Displacement [mm]

Figure 3.7: Effects over the loops by varying the relationship of β and γ

Gamma = 0.1, Beta = 0.9 Gamma = 0.5, Beta = 0.5 Gamma = 0.9, Beta = 0.1

25

20

15

10
Force [KN]

-5

-10

-15

-20

-25
-50.2 -50 -49.8 -49.6 -49.4 -49.2
Displacement [mm]

Figure 3.8: Detailed effects over the loops by varying the relationship of β and γ

When β > γ the hysteresis loops take the form of an “s”, which has not been observed in
the experimental test and goes against Equation 2.51. On the other hand, for the particular
case of γ = β = 0.5 the unloading branches are straight lines with a slope equal to k0 , as the
unloading branches of the bilinear model [59]. Hence, it was decided to set γ = β = 0.5.

The experimental data shows smooth transitions between the elastic and post-elastic part of
the force-displacement curve, see Figure 3.4. Figure 3.9 and 3.10 show how the response of
the Bouc-Wen model tends to that of the bilinear model for higher values of n. Thus, it was
decided to use n = 5.

36
n=1 n = 2.5 n=5
30

20

10
Force [KN]

-10

-20

-30
-50 -40 -30 -20 -10 0 10 20 30 40 50
Displacement [mm]

Figure 3.9: Effects over the loops by varying n

n=1 n = 2.5 n=5

20
Force [KN]

15

10

0
-0.5 0 0.5 1 1.5 2
Displacement [mm]

Figure 3.10: Detailed effects over the loops by varying n

Choice of parameters

Table 3.2 shows the values of the inputs used for the different Bouc-Wen models.

Table 3.2: Parameters of the Bouc-Wen model used in the implementation

Model k0 [N/m] a n γ β
Dolce’s test 3.39 1010 1 10−12 5 0.5 0.5
Bearing test 9.7 107 1 10−12 5 0.5 0.5

3.3 Generalization of the Bouc-Wen model

In order to improve the analysis and modelling of the bearing’s dynamic behavior, it was decided
to generalize the Bouc-Wen model. For this reason, The modelling of the hysteretic behavior of

37
the contact PTFE-steel is based on the proposed generalization of the Bouc-Wen model. This
proposal includes the effects of the sliding speed and contact pressure.

The parameter D determines the family of backbone curves that can be produced by the Bouc-
Wen model. Complementary, if the initial stiffness (k0 ) is assumed to be constant and the
ratio between the initial and post-initial stiffness (a) tends to zero, as the experimental data
both from the bearing test and the linear sliding test suggests, the plastic limit displacement
(D) has to describe the force-displacement relationship and, as a result of that, the coefficient
of friction. Therefore, these parameters are related by the following equation where Fp is the
plastic force:

Fp = k0 D (3.13)

On the other hand, as mentioned in Section 2.1, the friction force is determined by the following
equation:

F = µ N = µ (v, p) p A (3.14)

Where N is the normal force, µ is the friction coefficient which depends on the pressure (p) and
the slip rate (v) as explained above in Section 3.1 and A is the contact surface between both
sliding bodies. Thus, D can be determined by the following relationship:

µ (v, p) p A
D= (3.15)
k0

This will lead to a Bouc-Wen model restoring force:

F (t) = a k0 u(t) + (1 − a) µ p A z(t) (3.16)

Comparison between the classical Bouc-Wen model and its generalization is given in Section
5.1.1.

3.4 SDOF models

Single degree of freedom models based on the generalization of the Bouc-Wen model were
created to simulate both the linear sliding device experiment and the bearing test.

The parameters of the models were defined according to the limitations mentioned above and
within the range of those limitations, the value that was finally used was determined in order
to be consistent with the physical effects and the experimental data. The Bouc-Wen model
parameters are given in Table 3.2.

38
F,u
k

Figure 3.11: Sketch of the Bouc-Wen user element

3.4.1 Matlab

The first and simplest model was developed in Matlab implementing the Bouc-wen model.
Since both experiments are “displacement controlled”, the only input required to modelize
the behaviour that it is not known since the beginning would be the hysteretic value z. The
differential equation that governs z, Equation 2.45, was solved using the “classical Runge-Kutta
method” or “RK4”, which is a widely used mathematical method. The reason is that, for a
given step size increment h, the total accumulation of error is in the order of h4 . A Matlab
function was developed according to what was explain in Section 3.1 to determine the value of
D at each increment.

3.4.2 User-element in ABAQUS

One of the main FE-codes used in civil engineering, ABAQUS, permits to develop user-defined
elements by writing Fortran subroutines in combination with the standard libraries of ABAQUS.
A user element subroutine defined in [3] was modified in order to implement the modification
of the Bouc-Wen model according to Section 3.3.

The user-element implementation is limited to transient dynamics and the trapezoidal rule
without numerical dissipation was used to solve the equations of motion, this is an implicit
numerical method solver of differential equations. The solution of the hysteretic variable uses
the implicit Newton’s method.

Description

The user-element, see Figure 3.11, is built by two nodes connected by a linear spring, a dashpot
and the Bouc-Wen element, see Section 2.4. Those elements were placed at all the six degrees
of freedom (three displacements and three rotations). The macro elements considered are a
linear spring, a regular dashpot and a non-linear Bouc-Wen element, where the modifications
were introduced.

Note that any of these macro elements can be activated or deactivated at any direction de-
pending on which kind of analysis is going to take place.

The implementation required introducing new functions in order to deal with D varying over

39
the time due to the velocity and the pressure. In the same sense as Matlab, a function that
introduces the effect of the velocity over the friction coefficient was introduced. As the element
aspires to represent a general case, the hysteresis can be produced at any of the six degrees of
freedom.

In order to calculate both the normal force (FN ) and the contact pressure - the normal force
divided by the tangential area (AT ) - it has been required to use the equation of motion,
accounting the effect of the spring (kN ), the dashpot (CN ) and the mass (mN ) in the normal
direction, see the following equation:

FN (t) = mN u¨N (t) + kN u(t)N + CN u˙N (t) (3.17)

It could be argued that the magnitude of the force will be mainly influenced by elastic part
of the normal force (kN u(t)N ) making the other sources of the normal force almost negligible;
although, the results obtained at any step will determine the analysis made by ABAQUS at
the following one. Thus, this approximation will provoke problems especially for more complex
analysis, such as bridge implementation, where the inaccuracy between each step might be large
enough to prevent ABAQUS of finalising the analysis or to obtain a rather wrong solution. For
this reason, this approximation has been avoided.

Consequently, the degrees of freedom are uncoupled where there is no hysteresis, however,
where the hysteresis appears the tangential directions depend on their normal counterpart, see
the following equation:

FN (t) µ (u˙T (t), FANT(t) )


FT (t) = (3.18)
AT

The friction forces in the tangential direction (FT (t)) depend on the normal force (FN (t)) and
the friction coefficient (µ), which also depends on the contact pressure ( FANT(t) ) and the sliding
velocity (u˙T ) as explained before.

Particular case: displacement controlled experiment

The external effects that will be applied to the top node are: a concentrated load and a forced
displacement. The load must be consistent with the expected pressures at the contact surface
established in each experiment. This requires adjusting both the force and the contact area.
The forced displacement is a sinusoidal wave and it will have a certain amplitude.

The hysteresis in this particular case will only take place over the contact PTFE-steel and only
in one of the two possible perpendicular directions to the normal force in order to be consistent
with the experiments. All of these constraints make the problem bidimensional and only the
vertical and horizontal directions have been considered.

In addition, it is important to remark that the fixed boundaries at the bottom node are just
the degrees of freedom of the displacements. The degrees of freedom associated with all the
rotations have been set free.

40
Figure 3.12: Sketch of the Supports (from [2])

Bearing’s rotational degrees of freedom

The rotational degrees of freedom have a measurable relationship with the horizontal deforma-
tions that occur at the bearing due to the geometrical disposition, see the Sketch (Figure 3.12)
and the following equation:

u(t)T = h tg(θ(t)) ≈ h θ(t) (3.19)

Where u(t)T is the horizontal deformation of the bearing, h is the diaphragm height and θ(t) is
the angle between the non-deformed shape of the column (vertical direction) and the deformed
shape of the column. Due to the small deformations the tg(θ(t)) tends to θ(t).

However, this relationship has not been implemented in the current version of the subroutine
because it has been assumed to be negligible.

3.5 ABAQUS: 3D implementation

During the previous chapters it was introduced that some of the issues related to this thesis
will be modelled with the FE-code ABAQUS that unlike the other models used in the thesis, it
introduces several thousands of degrees of freedom. The limitations of FE-codes are well-known
for their regular users. Time-consuming and convergence are the main issues that any user
have to deal with, too sophisticated and elaborated models are usually not the best solutions
to address the problem since they take too much computational time and also convergence
problems. However, simplifying the problem by focusing on the most important parts and
diminishing non-important details will save a lot of time and even giving better solutions.
These could be applied by choosing proper meshes, applying symmetries or replacing parts of
the test by boundary conditions for instance. All these ”tricks” are commonly used and of
course they will be applied during this study.

41
3.5.1 Linear sliding device test implementation

The direct implementation of the PTFE-Steel contact in a FEM-model of a bearing introduced


too much unknowns and variables to reach convergence, specially due to the lack of experimental
data to compare. This problem has been solved by implementing a representation of the Linear
sliding device test (Dolce’s test) by a FEM-Model in ABAQUS.

There are only two kind of materials involved in this simulation and they will be defined by
their density and their elastic and damping properties. This values are commonly used and
they can be found and contrasted by different sources relatively ease.

Table 3.3: Material parameters in Dolce’s implementation

Material Density [kg/m3 ] Young's Modulus [Mpa] Poisson's Ratio α β


Steel 7850 210000 0.3 0.46 4.6−5
PTFE 2200 5000 0.46 1.39 3.71−3

Different models have been implemented with different ways of introducing the contact steel-
PTFE. From those, two of them are finally shown in this chapter in order to show the importance
of the contact area and the definition of the dimples. In addition this two models will be tested
by defining the contact steel-PTFE only by a sliding velocity dependency for the medium
contact pressure tested by Dolce et al. in the pads which had an average of 18.7 Mpa or by a
sliding velocity and pressure dependency (the average will be 18.7 Mpa but of course it is not
equally distributed in the face of the pads). The implementation of the friction coefficients has
been finally implemented by the Penalty method (see Section 2.1.3).

1. Equivalent area.
This first simplified model introduces one PTFE pad in each side of the steel (instead of
the 2 PTFE pads in Dolce's experiment) with a contact area equivalent to the area used
in the experiment. In addition the boundary conditions and forces are directly applied
to the PTFE pads diminishing the effect of the steel sheets of the outer sides. It can be
seen in Figure 3.13.

2. Modelling the dimples


This second model tries to be more accurate and realistic. The 4 PTFE pads are modelled
(see Figure 3.14) and an equivalent dimple pattern is introduced in the PTFE pads (see
Figure 3.15). In addition the effect that the steel sheets from the outer sides could have
in the contact is introduced by two steel shells.

3.5.2 Bearing test implementation

The 3D implementation of the laboratory test performed by M.Ülker-Kaustell is together with


the Bouc-Wen model, the main objective of this research.

The following materials were used to model the bearing test in ABAQUS. Their elastic prop-
erties can be seen in Tab 3.4.

42
Figure 3.13: Simplified model for Dolce's experiment

Figure 3.14: Model with dimples in the PTFE pads for Dolce's experiment

43
Figure 3.15: Detail of the dimples introduced in the PTFE pad

• Steel
• Copper
• PTFE
• Rubber

ABAQUS provides several options to simulate an hyperelastic isotropic material as the rubber
by incorporating the strain-energy function. The strain-energy function will be calculated
in terms of principal stretches and principal stresses in the FE-solution of boundary-value
problems. The selection of the parameters to define the strain-energy function depends on each
particular rubber which makes it a crucial and difficult choice in this study, especially taking
into account the current impossibility to test experimentally the rubber.

For computational analysis, the rubber will be considered incompressible so the constraint from
Equation 2.43 will be considered. As Holzapfel [42] recalls in his book, a “ very sophisticated
development for simulating incompressible materials” is due to [44] where the principal stretches
changes from the reference to the current configuration are defined as:

N
X µp
Ψ = Ψ(λ1 , λ2 , λ3 ) = (λαp αp αp
1 + λ2 + λ3 − 3) (3.20)
p=1
αp

Which comparing with the linear theory, the next condition is obtained:

N
X
2µ µp α p with µp αp > 0, p = 1, ..., N (3.21)
p=1

Being µ the classical shear modulus, the N the numbers of terms in the strain energy function,
and α a dimensionless constant.

44
Therefore a particularization of this model will be used for the model of the rubber in ABAQUS.
It is called the Mooney-Rivlin model, which was presented by M. Mooney [49] and R. S. Rivlin
[50]. This model sets N = 2, α1 = 2, α2 = −2. Using the definition of the invariants:

I1 (b) = trb = λ21 + λ22 + λ23 (3.22)

1
I2 (b) = [(trb)2 − tr(b2 )] = trb)−t detb = λ21 λ22 + λ21 λ23 + λ22 λ23 (3.23)
2

I3 (b) = trb = λ21 + λ22 + λ23 (3.24)

The expression Eq (3.20) can be rewritten as:

Ψ = c1 (λ21 + λ22 + λ23 − 3) + c2 (λ−2


1 + λ−2
2 + λ−2
3 − 3)
(3.25)
= c1 (I1 − 3) + c2 (I2 − 3)

where c1 = µ1 /2 and c2 = −µ2 /2 and therefore the shear modulus can be determined as
µ = µ1 − µ2 .

Finally the coefficients are related to the shear modulus recommended by the mentioned Pot
bearings [12] Euro-norm and based in the relation given by Anand [51] for a Mooney-Rivlin
strain energy potential that are c1 = 0.4375µ and c1 /c2 = 7 are expressed in Table 3.5

Table 3.4: Material parameters in the bearing implementation

Material Density [kg/m3 ] E [Mpa] ν α β


Steel 7850 210000 0.3 0.46 4.6eE-5
PTFE 2200 5000 0.46 1.39 3.71E-3
Brass 8490 100000 0.3 1.39 3.71E-3
Rubber 1230 - - 1.39 3.71E-3

Table 3.5: Rubber hyperelastic parameters

Material Shear modulus [Mpa] C10 C01 D1


Rubber 0.9 400000 40000 0

It is also important to explain the friction that are produced between the different materi-
als. These are modelled by defining the coefficient of friction between the materials with the
exception of the PTFE-steel contact as it has been explained previously. Once again, these
coefficients of friction have been determined by contrasting different sources for lubricated con-
tacts and they are shown in Table 3.6. Further discussion of the parameters will be carried out
in the next chapter.

45
Table 3.6: Coefficients of friction in the interactions

Contact Coefficient of friction (µ)


Steel-Steel 0.1
Rubber-Steel 0.1
Brass-Steel 0.2

The representation of the whole test, including frame, eccentric cylinder to apply the load
and the cylinder to control the motion were discarded from the very beginning. They would
have added many elements that would not improved the model. Instead of them, boundary
conditions and forces are applied in ABAQUS. The bottom plate is fixed, while the vertical
force is applied by a concentrated force and the control displacement is applied to the sliding
plate in the top. In addition at a first stage the curvature of the piston and the clearance of
the rubber where diminished.

Different stages have been performed as knowledge and intuition around the problem improves.
The next Figure 3.16 shows the first approximation to the problem performed by the authors
of this work. It tries to describe with fidelity the laboratory test taking into account the
simplifications above mentioned.

However different tests performed with this model enforced the authors to develop a different
model, both for practical and essential issues (see section 5.2.2). Therefore the next model was
proposed (see Figure 3.17) where the pot and sliding plate are much reduced in terms of size
and thickness, also where the PTFE-plate has been replaced only by its effect in the contact
with the steel and where the curvature of the piston has been introduced while it has been
made a hollow element with a thin layer of elements.

This model allows faster calculations while more accurate results are obtained due to the curve
implementation.

Sliding Plate
Piston

PTFE
Sealing ring
Materials
- Steel
- PTFE
- Brass
- Rubber Rubber
Pot

Figure 3.16: First bearing model cut in half by a vertical plane

46
Sliding Plate
Piston

Interaction PTFE-steel

Materials Sealing ring


- Steel
- Brass
- Rubber Pot Rubber

Figure 3.17: Simplified bearing model cut in half by a vertical plane

47
Chapter 4

Results

This chapter presents the most relevant results obtained during the research. Section 4.1 shows
the different tests carried out for the linear sliding device while Section 4.2 does it for the
laboratory bearing test

4.1 Linear sliding device

4.1.1 Non-lubricated contact

This section contains an SDOF analysis with two procedures, a Matlab code and a user element
subroutine and its later comparison.

Matlab

Figures 4.1 and 4.2 show the results obtained with the Matlab model.

15

10

5
Force [kN]

-5

-10

-15
-50 -40 -30 -20 -10 0 10 20 30 40 50
Displacement [mm]

0.1 1
Friction coefficient [-]

0.08
0.5

0.06
z [-]

0
0.04

-0.5
0.02

0 -1
0 5 10 15 20 25 30 0 5 10 15 20 25 30
time [s] time [s]

Figure 4.1: MatLab model of the modified Bouc-Wen model for a velocity of 15 mm/s with
non-lubricated contact

48
30

20

10

Force [kN]
0

-10

-20

-30
-50 -40 -30 -20 -10 0 10 20 30 40 50
Displacement [mm]

0.2 1
Friction coefficient [-]

0.15 0.5

z [-]
0.1 0

0.05 -0.5

0 -1
0 0.5 1 1.5 2 0 0.5 1 1.5 2
time [s] time [s]

Figure 4.2: Matlab model of the modified Bouc-Wen model for a velocity of 316 mm/s with
non-lubricated contact

6 Velocity = 15 mm/s
Velocity = 316 mm/s

2
Force [KN]

-2

-4

-6
-60 -40 -20 0 20 40 60
Displacement [mm]

Figure 4.3: User element modelization of the modified Bouc-Wen model for both 15 mm/s and
316 mm/s with non-lubricated contact

At the same time, it is shown how the friction coefficient and the hysteretic parameter z vary
during the analysis. Note that the variation of the coefficient of friction is quite significant and
clearly much higher in the case of high velocity.

User element

Figure 4.3 shows a the results obtained with the user element for both velocities.

Comparison SDOF non-lubricated

The following Figures 4.4 and 4.5 show a comparison between both SDOF models and the
experimental data. It can be seen that both from a qualitative and quantitative both models
reproduce rather efficiently the actual results.

49
4 Experimental data
BW Matlab
BW User element
3

2
Force [KN]

-1

-2

-3

-4

-50 -40 -30 -20 -10 0 10 20 30 40 50


Displacement [mm]

Figure 4.4: Comparison between SDOF models and experimental data for 15 mm/s with non-
lubricated contact

Experimental data
6 BW Matlab
BW User element

2
Force [KN]

-2

-4

-6
-60 -40 -20 0 20 40 60
Displacement [mm]

Figure 4.5: Comparison between SDOF models and experimental data for 316 mm/s with
non-lubricated contact

Moreover, both SDOF models generate, essentially, almost the same result for both velocities,
the difference is even more negligible in the case of higher velocities. Note that both of them
are built on the same behavioral equations. In the case of 15 mm the difference between both
models is less than 2 % regarding the maximum friction force and in the case of 316 mm the
divergence is around 0.06 %.

4.1.2 Lubricated contact

The folowing Figures, 4.6 and 4.7, show the results obtained with the SDOF model in Matlab for
the lubricated surface between PTFE-steel. With the Lubricated contact the difference between
both SDOF, the one performed in Matlab and the one produced by the user element subroutine
is lower than 0.06 %. As it can be appreciated the friction coefficients are significantly lower
than the non-lubricated case.

50
1

0.5
Force [KN]

-0.5

-1
-50 -40 -30 -20 -10 0 10 20 30 40 50
Displacement [mm]

0.02 1
Fricition coefficient [-]

0.015 0.5

z [-]
0.01 0

0.005 -0.5

0 -1
0 50 100 150 200 0 50 100 150 200
time [s] time [s]

Figure 4.6: SDOF modelization of the modified Bouc-Wen model for a velocity of 15 mm/s
with lubricated contact

0.5
Force [kN]

-0.5

-1
-50 -40 -30 -20 -10 0 10 20 30 40 50
Displacement [mm]

0.025 1
Friction coefficient [-]

0.02
0.5

0.015
z [-]

0
0.01

-0.5
0.005

0 -1
0 0.5 1 1.5 2 0 0.5 1 1.5 2
time [s] time [s]

Figure 4.7: SDOF modelization of the modified Bouc-Wen model for a velocity of 316 mm/s
with lubricated contact

In both figures the hysteretic loop can be seen for both extreme velocities. The most interesting
thing regarding the lubricated contact at Dolce’s experiment is that the values of the friction
coefficient do not have a significant variation from the lowest to the highest velocity.

51
4.1.3 ABAQUS: 3D Implementation

Previous sections have shown the results obtained by ABAQUS in a general case for Dolce’s ex-
periment. The next figures will show the results obtained by the two different models proposed
to simulate with finite elements the experiment and see the relevancy of the degree of detail.
Once again, this analysis has been performed for the contact pressure of 18.7 Mpa and T = 20
ºC for the lubricated and non-lubricated and two different velocities (15 and 316 mm/s) so it
can be compared to Dolce 's results.

In order not to overdo the graphs, the next correspondence will be used:

• Model 1: Model with dimples, velocity dependency for a fixed pressure 18.7 Mpa.
• Model 2: Model with dimples, velocity and pressure dependency.
• Model 3: Equivalent area model, velocity dependency for a fixed pressure 18.7 Mpa.
• Model 4: Equivalent area model, velocity and pressure dependency.
• Model 5: Experimental data extracted from Dolce 's paper [10].

Figure 4.8: Non-lubricated model for the small velocity 15 mm/s

Figure 4.9: Lubricated model for the small velocity 15 mm/s

52
Figure 4.10: Non-lubricated model for the high velocity 316 mm/s

Figure 4.11: Lubricated model for the high velocity 316 mm/s

Probably the first thing the reader will notice is the big difference of accuracy of the models
between low and high speeds. While in low velocities the model gives a good approach both
qualitatively and quantitatively (relative error of 1.9 % in the non-lubricated case and 7.4 % in
the lubricated case, see Figures 4.8 and 4.9), for high velocities the model does not match so
precisely Dolce's results quantitatively (relative error of 17.8 % in the non-lubricated case and
46.1 % in the lubricated case, see Figures 4.10 and 4.11) and even not qualitatively when the
contact pressure is not considered as it can be seen in model 1 and model 3 in Figure 4.11.

Non-consideration of the pressure usually has a greater effect on the lubricated model for high
velocities, because the pressure is taken as bigger as it should be due to the pressure distribution
gradient that occurs in the contact PTFE-steel. Figure 4.12 shows an example of this pressure
distribution.

53
Figure 4.12: Contact pressure in the contact PTFE-steel before the movement of the steel
occurs in the detailed model with dimples

4.2 Bearing

4.2.1 Multidirectional pot bearing

The following Figure 4.13 shows how both models, 3D in ABAQUS and the SDOF model, both
in Matlab and in ABAQUS with the user element subroutine, behave. It can be seen that there
is a rather good agreement for higher pressures and frequencies, i.e. sliding velocities. However,
for low pressures and sliding velocities the models are not accurate, as expected. In addition,
it can be seen that the results produced by both kinds of models (3D and SDOF) are quite
similar, almost the same.

Due to the bottom limit constrain explained in Subsection 3.1.2 the SDOF models were adjusted
according to the experimental data from KTH. Figure 4.14 shows the result. It can be seen
that a quite good agreement was obtained with the SDOF models.

54
Figure 4.13: Comparison between experimental data and bearing’s model in ABAQUS (3D)
and SDOF model using the lubricated model defined by Dolce

55
Figure 4.14: Comparison between experimental data and bearing’s model in ABAQUS (3D)
using the lubricated model defined by Dolce and SDOF model for the multidirectional bearing

56
57
Chapter 5

Discussion

This chapter summarizes and discusses the information obtained within this research and try
to obtain conclusions and ideas that will help further research.

In Section 5.1 we give a comparison between the classical and generalized Bouc-Wen models
used to model pot bearings. Section 5.2 gives a discussion of the 3D FE-models and how we
believe that they could be improved. A comparison between the Bouc-Wen models and the 3D
is given in Section 5.3 and then, Section 5.4 aims at describing the thermomechanical aspects
of steel-PTFE contacts which have been neglected in the models studied in this thesis. Section
5.6 discusses the subject of a uni-directional pot bearing. Finally Section 5.7 summarizes the
conclusions of the thesis and it also suggests future researrch within the topic of the study.

5.1 SDOF models

5.1.1 Comparison between the classical and generalized Bouc-Wen


models

As it was explained in Section 3.2, it was decided to choose a set of Bouc-Wen classical pa-
rameters that would produce a response tending to the bilinear model. Their effect over the
loop in the classical definition of model was studied. This Section will discuss the effect of the
parameters regarding the generalized Bouc-Wen model as well as a comparison between both
simulations and the experimental data.

Figure 5.1 shows the effect over the loop that the ratio between the elastic and post-elastic
stiffness has.

58
a = 1e-3 a = 1e-4 a = 1e-5 a = 1e-12
30

20

10
Force [KN]

-10

-20

-30
-50 -40 -30 -20 -10 0 10 20 30 40 50
Displacement [mm]

Figure 5.1: Effects over the loops by varying a

It can be appreciated that the effect is quite similar both qualitatively and quantitatively in
relation to the classical Bouc-Wen model, see Figure 3.5. Once again, The effect over the loop
is quite negligible when a ≥ 10−4 .

γ and β

Figures 5.2 and 5.3 show how the loop varies for different set of parameters.

Gamma = 0.1, Beta = 0.9 Gamma = 0.5, Beta = 0.5 Gamma = 0.9, Beta = 0.1
30

20

10
Force [KN]

-10

-20

-30
-50 -40 -30 -20 -10 0 10 20 30 40 50
Displacement [mm]

Figure 5.2: Effects over the loops by varying the relationship of β and γ

Gamma = 0.1, Beta = 0.9 Gamma = 0.5, Beta = 0.5 Gamma = 0.9, Beta = 0.1

15

10

5
Force [KN]

-5

-10

-15

-20
-50 -49.95 -49.9 -49.85 -49.8 -49.75 -49.7 -49.65 -49.6
Displacement [mm]

Figure 5.3: Detailed effects over the loops by varying the relationship of β and γ

59
n=1 n = 2.5 n=5 n=8
30

20

10
Force [KN]

-10

-20

-30
-50 -40 -30 -20 -10 0 10 20 30 40 50
Displacement [mm]

Figure 5.4: Effects over the loops by varying n

n=1 n = 2.5 n=5 n=8


25

20
Force [KN]

15

10

0
-0.5 0 0.5 1 1.5 2
Displacement [mm]

Figure 5.5: Detailed effects over the loops by varying n

From a qualitatively point of view the “s” shape produced when β is higher than γ seems to
have been softened.

Figures 5.4 and 5.5 show how n affects the transition between elastic and post elastic.

It can be seen that the response is more sensitive to this parameter at the beginning of the
curve, although, when the cycles starts the influence is reduced quite much, which, in contrast,
did not happen with the classical Bouc-Wen model, see Figure 3.9.

In general, it can be seen that the response is less sensitive to the choice of parameters since
the inclusion of D as a function of the actual forces and stiffness of the system seems to be
more influential. Moreover, Figure 5.6 shows a comparison between the experimental data and
the classical and the generalized versions of the Bouc-Wen model.

60
Experimental data Generalized Bouc-Wen model Classical Bouc-Wen model
6

2
Force [KN]

-2

-4

-6
-60 -40 -20 0 20 40 60
Displacement [mm]

Figure 5.6: Comparison between the experimental data, the classical and the generalized Bouc-
Wen models

The generalized model is able to reproduce the experimental data more accurately both from
a qualitatively and quantitatively point (Note that the simulations are not producing the same
maximum displacement because they are producing the theoretical one of 50 mm, which is
slightly smaller that the actual value measured).

As stated in Section 2.4.2 D can be seen as the yield displacement of the corresponding bilinear
model and, in combination with k0 , a and n, establishes the family of backbone curves that can
be produced by the classical Bouc-Wen model. Therefore, the response is quite constrained by
this parameter and it makes quite difficult to generalize the use of the Bouc-Wen model or to
forecast responses. This can be even more complex when non-linear materials are considered,
because their yield displacement might vary significantly depending on the particular conditions,
requiring a different value of D according to the different scenarios.

Consequently, the generalization means an improvement of the model response, since it is


able to account for those differences through the relationship between D and the actual forces
acting in the system at any specific moment. Therefore, the model is not only able to improve
the simulation of the non-linearities of the materials but also to vary the response due to
any alterations of the particular conditions of the system, in terms of variations of force, i.e.
pressure, and slip rate.

Thus, the main enhancement of the present generalization is that it defines an infinite number
of backbones curves instead of just a family as the classical model does. That infinite number of
backbones curves depend on the actions applied to the system making the model more robust
and flexible.

Another important consideration is that, from a wider perspective, this generalization could be
theoretically applied to any other system with mechanical hysteresis. D is just being obtained
by taking into account the physical effects that are taking place, which in this particular case
are the friction forces, but this concept could be extended, at least in principle, to any other
type of structural system by obtaining an equation that relates the yield displacement with the
physical phenomena.

61
5.1.2 Sensitivity Analyses of the contact area for the sliding test
implementation

The purpose of this section is to analyze the influence of the contact area over the Matlab and
the user-element model and its comparison with Dolce’s lab test. Both SDOF models produce
almost the same response and at the same time from a qualitative perspective, they behave
quite similarly when they are compared with the experimental data.

However, when both models are compared with Dolce’s experimental test from a quantitative
point of view, larger discrepancies are obtained. Dolce’s paper does not specify the actual
contact area of the PTFE and it is unknown how many dimples are distributed over the surface
since the only given information is the general pattern of the PTFE sheets and not the real pads
that were finally used. In other words, how the PTFE was cut and placed was not specified.
The total area of the PTFE, not considering dimples, is 4000 mm2 ; this area will be referred to
as the nominal area in the rest of the document.

Contact Area

The contact area is the surface where the friction takes place. The friction force is proportional
to the normal force and the friction coefficient, the normal force is determined by the pressure
and the contact area. The pressure may not be equally distributed.

What can be assumed from the given pattern, see Figure 1.3, is that approximately, before
applying any load over the PTFE, there would be a 20 − 30 % (2800 − 3200 mm2 ) of area lost
due to the dimples. This number is based on the maximum and minimum number of dimples
regarding the specific pattern of the PTFE’s sheets.

Another thing to consider is the determination of the real contact area at the beginning of the
process due to the mechanical deformations of the PTFE once the load is applied. In [62] an
extensive study of the of the properties of the PTFE in compression was conducted. According
to that study, the mechanical properties showed to be strongly affected by the strain-rate and
temperature. The study also provided strain-stress experimental relationships for different
temperatures and strain rates as well as the experimental poisson ratios. At the same time, the
handbook of PTFE properties [63] shows its mechanical properties not only under compression
but also how the material behaves under other loading conditions.

The problems considered in this thesis consist of two different load steps: first, a stationary load
(applied pressure or the weight of the bridge) and second a transient load or excitation (forced
displacement or a train passing). Thus, the initial real contact area is just determined by the
first load, which does not produce friction since there is no tangential movement. In addition,
due to the lack of friction no heat will be generated by this initial load, so the temperature is
contact and equal to the air temperature.

Figure 5.7 shows the deformation of a general dimple in the normal direction [63] under certain
conditions considered in Dolce’s experiment (linear sliding device test): pressure of 18.7 Mpa
and 20 ºC of air temperature. It has been assumed a low strain rate. The shape of the dimple
was determined according to the pattern of the PTFE used in the experiment.

62
1.4
Dimple before applying the load
Dimple once load is applied
Evolution of Dimples [mm] 1.2

0.8

0.6

0.4

0.2

0
0 1 2 3 4 5 6 7 8
Contact Line [mm]

Figure 5.7: Deformation of the dimples due to a specific external load (Pressure of 18.7 Mpa
and temperature of 20 º C)

This will provoke, with this specific load, that the real contact area will be increased by 47 %
of the dimples area. However, it is important to state that the tangential deformation has not
been considered, its inclusion will provoke an increase of the area but an increase of friction as
well due to the fact that the friction force at the contact will oppose the deformation in this
direction.

Analysis

This sensitivity analysis has been conducted using the Matlab model for the highest and lowest
sliding speed tested by Dolce: 15 mm/s and 316 mm/s, with a constant contact pressure of
18.7 Mpa and a constant temperature of the air of 20 ºC. The results of this analysis are the
same for both the lubricated and the unlubricated case.

The measurement of the error is focused on two different outputs of the system: the maximum
force achieved and the dissipated energy. The relative error has been calculated by comparing
those values obtained by Matlab with the ones that can be calculated using the hysteretic loops
provided by Dolce. The inscribed area within the loops, which represents the dissipated energy
as explained in Section 2.4, was obtained with Simpson method. The following equation was
used to calculate the relative error, where Er is the relative error, Xe is the experimental value
and Xm is the value provided by the model based on the experimental data.

Xe − Xm
Er = (5.1)
Xe

It is quite important to state now that those values were not explicitly provided in the paper;
i.e. there is no table of values that defines the loops of hysteresis in an unambiguous form. For
this reason, the values were obtained from the experimental loops by using a Matlab function to
create vectors from figures. Additionally, there is a spread in the experimental data, specially
in the lubricated case, that will provoke divergences between the experimental results and the
model proposed by Dolce. According to the possible sources of deviations mentioned before, it
has been considered reasonable to have a 10 % of error’s range, see figures 5.8 and 5.9.

63
25
Error of the total energy disipated
20 Error of the maximun force

15
Relative error [%]

10

-5

-10

-15

-20

-25
2000 2100 2200 2300 2400 2500 2600 2700 2800 2900
Area of Contact [mm 2 ]

Figure 5.8: Relative error between experimental data and SDOF model based on Dolce’s model
of friction (v = 15 mm/s)

50
Error of the total energy disipated
Error of the maximun force
40
Relative error [%]

30

20

10

-10

-20
1800 2000 2200 2400 2600 2800 3000
Area of Contact [mm 2 ]

Figure 5.9: Relative error between experimental data and SDOF model based on Dolce’s model
of friction (v = 316 mm/s)

What can be seen from those figures is that the contact area required by the Bouc-Wen model
in order to be totally consistent with the experiments varies in the case of the maximum force
and in the case of the energy dissipated. The dissipated energy requires lower values of the
contact area to reduce the error than the force.

On the other hand, when it is studied how the area affects the model for different sliding speeds,
it can be seen that the required area varies between low and high sliding speed. For 316 mm/s
the contact area would be within the range of around 1950 mm2 and almost 2200 mm2 and for
15 mm/s the range goes from less than 2350 to 2550 mm2 . The area of contact not considering
the mechanical deformations was within the range of 2800 and 3200 mm2 .

Another interesting consideration is that non-uniform pressure distribution could also influence
the results, because the SDOF models consider an equally distributed pressure.

The fact that the higher the sliding speed the smaller the value of the required contact area
by the Bouc-Wen model is in concordance with what was explained in Section 5.4, because
the increase of heat produces the reduction of the shear strength of the PTFE as a whole.

64
Therefore, given that the shear strength is opposing the friction forces within the PTFE, this
effect will have a negative impact over the resistance of the material and as a result of that a
reduction in the friction coefficient.

It is very important to realize of the fact that these figures highlight the extremely good
agreement that exists from a qualitative point of view between the models and the experimental
data. The reason is that both lines are almost parallel; this means that for an increase or
decrease of the contact area the error of both the force and the dissipated energy varies with
almost the same rate.

Finally, the reason why the dissipated energy has a higher error than the maximum force for
both velocities is because in the case of the lowest sliding speed the experimental loops were not
exactly symmetrical. In the case of the highest sliding speed the experiment produced loops
with 55 mm displacement instead of the theoretical 50 mm, see Section 5.3.

Discussion regarding the contact area

The main problem is the necessity of using the area of contact as a parameter for the models.
As was explained in Section 2.2, the processes that take place at the contact produce heat and
wear and as a result of that, the geometry and the properties of the contact surface change
with the appearance of transfer films or damaged areas.

Additionally, there are some uncertainties regarding the experimental data used by the authors
of this thesis. The contact temperature was not measured during Dolce’s experiment, although,
since the test measured final results of the friction its influence is implicitly included in the
results, see Figure 5.10.

However, the temperature field is unknown, which means that the influence of the flash tem-
perature cannot be measured. This is a problem in order to generalize the model since the
results can be significantly affected by the specific thermal configuration of the experiment,
which10 might differ considerably from other particular cases. Note that the fluctuation of the
temperature at the contact could be computed using the model that was explained in Section
2.2.1. 9
8

6 Contact temperature
Air temperature
5

4 Friction force Friction force

3
Measured Friction force
2

0 Figure 5.10: Sketch of the sliding device test


0 5 10 15 20 25 30 35 40

65
Another possible source of uncertainty regarding the sliding device test is the low rate of change
of PTFE pads, 300 test were performed and the pads were changed three times. Those pads
were measured for different pressures, sliding speeds and temperatures and as a result of that
they could have been damaged affecting the final results. Note that aside from the general
mechanical effects that take place at the PTFE-steel contact, regarding the dimples of the
PTFE, the wear due to friction would produce a recession of dimples developing new flat areas
over the surface, as observed in Dolce’s experiment.

In order to deal with the problems regarding the contact area, it has been considered that a
reasonable and quite simple approach would be to introduce a factor that modifies the contact
surface according with the different phenomena.

This will require the inclusion of a factor that would vary the real contact area depending on
how intense is the variation of temperature. Remember from [36] that the sliding speed is the
main factor behind the increase of heat. This could perhaps be modeled by an exponential
factor: cA (v) = e−αA v , because the higher the sliding speed at the contact the smaller the area
demanded by the model.

By using this factor, the real contact area will be higher and closer to the theoretical one when
the sliding speed (υ) is lower and the temperature of the contact is closer to the air temperature.
On the other hand, when the sliding speed increases, so does the dissipation of energy as well
as the temperature of the contact, leading to a reduction of the contact area. This new factor
will require more tests at different velocities to calibrate the αA value. However, this factor
might require the inclusion of pressure dependency as well, which increases the complexity of
the definition and calibration.

Finally, it can be said that the surface considered for the friction is a very sensitive value that
would require of further investigations in order to be properly implemented in more advanced
models, including the different thermomechanical effects that are taking place.

5.2 ABAQUS: 3D implementation

5.2.1 PTFE-steel contact implementation

Clearly, the initial assumptions that the most important non-linear phenomena to consider
in a 3D simulation of a pot bearing are the rubber hyperelasticity and the steel-ptfe friction
coefficient, where not correct. Therefore, in this section, we summarize the efforts made to try
to improve the 3D model of the tested multi-directional bearing and give suggestions as to how
the simulation results could be improved.

It was confirmed the initial intuition of the importance of the pressure dependency at every
node of the 3D models, see Section 4.12..

Secondly, if the two models in terms of complexity are compared (equivalent area versus PTFE
patterned) it is found that not big differences are found. Prior to the analysis, it was assumed
that the more detailed model would produce better results but, even if sometimes they do, it
has been observed that it is what it is actually happening. In contrast, the equivalent area is
closer to the experimental data in the lubricated case. These results show that the model is
not able to capture the difference between the dimple-patterned surface and an equivalent area

66
that as it is discussed in Section 5.1.2, should exist. This could be explained by the following
reasons:

• The model does not reproduce accurately the mechanical behavior of the PTFE. Thus,
the modeled material is too rigid and as a result of that the deformation of the dimples
is not simulated accurately enough.

• The pattern of the PTFE has been simplified in the 3D model. The dimples are cylindrical
while they are spherical in the real case. In addition, as mentioned in Section 5.1.2, there
is a lack of detail in the exact way to cut the PTFE.

In addition to the above mentioned factors, the velocity plays a crucial role in the modelling.
Even if the authors expected some difficulties to match higher velocities, due to higher devi-
ations between the proposed model of Dolce et al. and the experimental data, it seems that
the implementation works almost perfect for small velocities while high velocities have bigger
discrepancies.

Finally it should be commented some expectable effects that the theory chapter mentioned and
the current ABAQUS models are unable to capture. These are the breakaway effect (sometimes
it seems to appear, but it is due to numerical errors of the FE-code), the slip-stick effect and
the cyclic effect occurred because of the temperature in the contact.

5.2.2 Bearing model

The first model of the bearing test implementation showed that:

• Since this study is not interested in the stresses produced in the different steel parts, it is
not needed a full development of the thickness of the pot, the interior of the piston or the
sliding plate. A less accurate representation of those elements saves a lot of computational
time (less nodes and consequently, less degrees of freedom) and improves productivity.
The simplification of both parts have been done by making them thinner and reducing
their volume by introducing non-material regions, i.e. cavities and rigid body constrains.

• The curvature and clearance of the piston and rubber are important to allow the bearing
to develop the different sources of friction.

5.3 Comparison between SDOF and 3D models for the


sliding device test

For the velocity of 15 mm (Figures 5.11 and 5.12), it can be seen that the agreement between
both models and the experimental data is rather good both from a qualitative and a quantitative
point of view for both the lubricated and non-lubricated case, besides, the models produce
similar results. One interesting thing to point out is the differences between the experimental
data and the MatLab model and the ABAQUS’s case regarding the initial stiffness (k0 ). In the
first two cases, the beginning of the hysteretic loop is almost vertical while in the last one there

67
is a certain inclination lower than the vertical one. However, this is most likely caused by the
time-step in ABAQUS.

At the same time, the 3D ABAQUS model has some divergence between each cycle, this is also
caused by the time-step in ABAQUS.

However, the main differences arise between the models and the experimental loops on the right
side. SDOF and 3D models (accounting the time-step effect) produce symmetric responses,
which make sense from a theoretical point of view given that the experiment is governed by a
sinusoidal displacement. In contrast, the tests produce a non-symmetrical response.

For the velocity of 316 mm (Figures 5.13 and 5.14), it can be seen that there is a better
agreement between the SDOF model than in the case of the 3D ABAQUS model, both from
a qualitative and quantitative point of view. However, the SDOF model is defined in such a
way that it can be adapted for the different cases varying the considered area of contact, see
the sensitivity analysis performed in Section 5.1.2.

Therefore, the considered area has been varied depending on the degree of agreement with
Dolce’s hysteretic loops. Almost certainly, the effects that were mentioned in the sensitivity
analysis as a reason for the necessity of varying the contact area in the MatLab model are
provoking the divergence of the 3D ABAQUS model. For instance, effects such as the increase
of heat due to a higher velocity that reduces the friction coefficient.

In this case, the experimental loops are symmetrical as well as both models. The most significant
deviation between the Bouc-Wen models and the tests’ loops for the non-lubricated interaction
is the maximum displacement obtained. The reason behind this is that in Dolce’s paper is
stated that the maximum displacement was forced to be 50 mm but it can be appreciated that
in fact the maximum displacement is higher, around 55 mm. The models’ loops were set to
produce a maximum displacement of 50 mm.

4000
Experimental data SDOF 3D
3000

2000

1000
Force [N]

-1000

-2000

-3000

-4000
-60 -40 -20 0 20 40 60
Displacement [mm]

Figure 5.11: Comparison between the experimental data, the SDOF and the 3D models for a
sliding speed of 15 mm/s with non-lubricated surface

68
1 Experimental data SDOF model 3D model

0.5
Force [KN]

-0.5

-1

-60 -40 -20 0 20 40 60


Displacement [mm]

Figure 5.12: Comparison between the experimental data, the SDOF and the 3D models for a
sliding speed of 15 mm/s with lubricated surface

8000
Experimental data SDOF 3D
6000

4000

2000
Force [N]

-2000

-4000

-6000

-8000
-60 -40 -20 0 20 40 60
Displacement [mm]

Figure 5.13: Comparison between the experimental data, the SDOF and the 3D models for a
sliding speed of 316 mm/s with non-lubricated surface

1.5 Experimental data SDOF model 3D model

0.5
Force [KN]

-0.5

-1

-1.5

-60 -40 -20 0 20 40 60


Displacement [mm]

Figure 5.14: Comparison between the experimental data, the SDOF and the 3D models for a
sliding speed of 316 mm/s with lubricated surface

69
5.4 Thermomechanical processes PTFE-steel contact

The increase of contact temperature changes the PTFE structure. The elastic properties of the
PTFE as well as the shear strength are quite sensitive to the temperature, see Figure 5.15 and
Figure 5.16. This effect is even more important at the contact, where adhesive layers appear
between the PTFE and the steel provoking an increase of the friction coefficient.

Figure 5.15: Stress (Compression) – Strain relationship of the PTFE

70
Figure 5.16: Stress (Shear) – Strain relationship of the PTFE

There is a correlation between the sliding speed and the strain rates. Therefore an increase of
the sliding velocity provokes a rise of the shear of the PTFE which leads to an increase of the
friction forces[36]. On the other hand, an increment of the applied pressure produces an increase
of the amount of contact points between the asperities of the steel and the PTFE, i.e. there
is a rise of the contact area, which in general increases the temperature at the contact as well
because the PTFE has difficulties to dissipate heat due to its low internal heat conductivity.
Thus the easiest way to dissipate heat is through the boundaries of the PTFE that are in
contact with the atmosphere, so the longer the distance to those boundaries the less effective
is the dissipation of heat.

The increment of temperature at the contact PTFE-steel produces two opposite mechanical
effects. First, the shear strength of the PTFE as a whole is reduced and so are both the
coefficient of friction and the general elastic modulus of the PTFE. The shear loads tend to
produce failure along a plane parallel to their direction, in this case the shear strength of the
PTFE is opposing the friction forces at the contact PTFE-steel and the force that holds the
layer of PTFE. Therefore, if there is a decrease of the shear strength, given a constant value
of the normal contact force, there is a reduction of the friction forces at the contact. Second,
the reduction of the elastic modulus favors the formation of adhesive layers (transfer films)
composed of melted parts of the PTFE at the asperities of the steel; they increase the contact
area and the friction coefficient. Thus, the sum of both effects would determine how the friction
behaves. Note that previous studies ([36, 10]) showed that the increase of the pressure produces
a decrease of the friction coefficient. The reason seems to be that the decrease of the shear
strength of the PTFE as a whole has a more significant influence [36].

On the other hand, the flash temperature or the temperature variations of the PTFE are
determined by two opposed effects as well. The transfer films are softer than the non-damaged
PTFE surface, where a much more severe contact is produced. In addition, they reduce the
temperature at the contact by absorbing heat. Nevertheless, the capacity of the rest of the
PTFE to dissipate heat is quite low, so, at the end, the transfer films are forced to gather heat.
Hence, the final variation is determined by summing up both effects.

71
The PTFE properties are highly determined by the specific conditions at the contact. For this
reason, the knowledge of the temperature field within PTFE is needed in order to calculate the
stress and the friction field accurately [34].

5.5 Comparison between 3D and SDOF models for the


laboratory test

As it can be seen in Figures 4.13 and 5.19, the results obtained in both cases are quite similar,
which suggests that from a quantitative point of view, the behavior is mainly determined by
the PTFE-steel interaction. Hence, this implies that some fudamental mechanisms are not
accounted or are not relevant in the 3D models. These could be:

1. Interactions rubber-steel and brass-steel

2. Hyperelastic behaviour of the rubber

However, some differences arise regarding the qualitative perspective. The SDOF models are
not able to reproduce the negative slope that appears in some cases in the post-yielding part
of the hysteresis loop while the 3D models can. This variation is probably produced by the
rubber-interaction as it was previously explained in Section 5.2.2. Although, this could be
achieved by introducing another macro element, transforming the SDOF into a MDOF model
with two macro elements.

5.6 Uni-directional pot bearing laboratory test

A unidirectional pot bearing was also tested. The difference between the unidirectional and
multidirectional pot bearing is that the former can only move in one direction in the horizontal
plane due to a guide which is placed inside the mechanism, the latter is free to move in both
directions of the horizontal plane. Remember that in both cases the surface contact of PTFE-
steel is lubricated.

5.6.1 Comparison with the multidirectional bearing

Figure 5.17 shows a comparison between the experimental results of both types of bearings.

As can be seen, there is more friction or dissipation of energy in the unidirectional bearing test
than in the multidirectional bearing. On the other hand, from a qualitatively point of view
the hysteretic loops produced by both experiments are quite similar. It is also interesting to
see how the stick-slip is much higher and influential in the case of the unidirectional bearing,
especially for the cases with the higher values of pressure and sliding speed. Complementary,
it should be noted that the initial stiffness is the same for both cases, the guide does not make
the contact stiffer and the break away phenomena is quite similar for both cases.

72
There is an increase of friction from the unidirectional to the multidirectional bearing in all the
cases. The increase of the maximum friction force has been measured by using the following
equation:

Fu − Fm
I= (5.2)
Fm

Where Fu is the maximun friction force at the unidirectional bearing and Fm its counterpart
at the multidirectional one. The range of increases goes from 25 % to 156 %. The average
increase is 70 %, having the maximum values for the lowest value of the sliding speed, being
around the 100 % increase for all the different pressures, see Table 5.1.

Table 5.1: Increase of friction (%)

Force Frequency = 0.01 Hz Frequency = 1 Hz Frequency = 2 Hz Frequency = 5 Hz


10 KN 130 % 25 % 39 % 56 %
50 KN 97 % 35 % 27 % 86 %
100 KN 137 % 32 % 38 % 81 %
200 KN 156 % 54 % 59 % 102 %

The main possible source of this increase of the friction forces is the interaction PTFE-guide.
Geometrical imperfections in the test rig generate an eccentricity that produces a moment
when applying the different testing loads over the device that puts in contact the guide with the
PTFE. Furthermore, those geometrical imperfections are not removed so the moment produced
has always the same tendency, i.e. the same sign. Hence, the contact area between the PTFE
and the guide is the same for the different experiments, producing an intense local wear. In
conclusion, PTFE and guide are in contact over an specific region during the whole experiment
and as a result of that an extra friction force appears generating another source of dissipation
of energy.

5.6.2 Comparison with the linear sliding device test

Figure 5.18 shows a comparison between the unidirectional bearing test and both the lubricated
and non-lubricated sliding device test.

It can be seen in the figure how there is a divergence between the lubricated experiment per-
formed within Dolce’s paper and the one performed with the unidirectional bearing. The
coefficients of friction obtained with the unidirectional bearing test appear between the ones
of the lubricated and non-lubricated sliding device test, being closer to the lubricated case for
higher pressures and closer to the non-lubricated case for lower pressures. This divergence pro-
voked the necessity to develop an adjustment of the lubricated models used for multidirectional
bearing and the lubricated model proposed by Dolce. The experimental data used to develop
the model was the test of the unidirectional pot bearing for an enforced 2 mm amplitude of
displacement.

73
Figure 5.17: Comparison between multidirectional and unidirectional bearing test

74
Unidirectional bearing test
Dolce's lubricated test
Dolce's Non-lubricated test

Fricition coefficient [-]


0.2

0.15

0.1

0.05

0
0
50 0
100 5
150 10
200 15
Velocity [mm/s] 250 20
300
350
25 Pressure [MPa]
30

Figure 5.18: Comparison between both sliding device lubricated and non-lubricated test and
unidirectional bearing test

5.6.3 Unidirectional pot bearing friction modelling

Due to the divergence that was found between Dolce’s proposal for the lubricated interaction
of PTFE-steel and the multidirectional bearing test with the unidirectional bearing test, it was
decided to develop a new model that would reproduce the test’s results for the particular case
of 2 mm displacement, as explained in Section 3.1.

Figure 5.19 shows the comparison between the model proposed for the unidirectional pot bearing
and the experimental data obtained in the lab. The model was based on the experimental data
obtained with the unidirectional bearing.

As it can be seen the agreement is quite good for almost all the cases, the main differences arise
for the higher values of the sliding speed and pressure. SDOF models and 3D model produce
almost the same response, as expected, implying a very strong influence of 3D model to the
PTFE surface interaction definition.

5.6.4 Agreement between the unidirectional bearing model and new


unidirectional bearing test data

Aside from the 2 mm displacement set, which was used to develop the unidirectional bearing
model mentioned in the previous Section, more tests were performed with the unidirectional
bearing at the laboratory. It is important to state now that the subsequent experiments were
not performed with the multidirectional bearing.

The other enforced amplitudes of displacement are 5 mm and 1 mm. Figures 5.20, 5.21 and
5.22 show the different values of the friction coefficient corresponding with different amplitudes
of displacement. The so called cyclic effects are also showed in the figures, indicating how the
friction decays due to this phenomenon.

From the figures it can be observed that there are significant fluctuations between each set of
experiments, reducing the accuracy of the model. However, it is important to highlight that
from a qualitative point of view the friction coefficient seems to have a similar tendency for all
the cases.

This thesis will not be able to explain why this divergence between different sets of experiment
is observed but there are some ideas about the reasons that should be exposed:

75
Figure 5.19: Comparison between experimental data and unidirectional bearing’s model in
ABAQUS (3D) and SDOF model of the multidirectional bearing

76
• Thermal effects: it has been extensively explained in Section 2.2 and 5.4 the strong
influence that the thermal effects have over the PTFE behavior, especially over the shear
strength. The temperature field of the contact area has not been calculated within this
thesis and obviously it will be different depending on the sliding velocities and contact
pressures, which could be a possible source for the fluctuations of the friction coefficient.

• Mechanical effects: they are strongly influenced by the thermal configuration as well,
as described in Section 5.4. In addition, the mechanical properties have not been fully
considered within this thesis. Under compressive loading the stress-strain relationship in
the normal direction is quite sensitive to the strain rate, i.e. how fast the mechanical
deformation is produced [62]. Hence, different values of the strain rate provoke that the
polymeric material can be more or less resistant to the same applied load.
If this idea is applied to the tangential direction, theoretically the shear strength of
the PTFE would vary depending on the strain rates, in other words, different sets of
experiments may produce different responses for the same values of contact pressure and
sliding speed because the strain rates would be different. Consequently, the PTFE would
be more or less robust depending on the specific conditions, producing higher or lower
friction forces.

• Lubrication: The effect of the lubricant over the PTFE-steel contact interface has not been
quantified. For instance, it has been observed that in the case of an earthquake the friction
coefficient between PTFE-steel at the bearings can vary from 0.10 and 0.15 depending on
the lubrication, which indicates how there is such a big uncertainty regarding this topic
[64].
It is also true that the model developed due to the sliding device test generally reproduce
quite well the friction force produced at the multidirectional bearings. Thus, the sensi-
tivity of the friction coefficient to the lubricant does not seem to be very decisive but it
could contribute to the fluctuation of the results.

In principle, all of these ideas could be applied to the multidirectional pot bearings.

0.14
1 mm with cyclic effect
1 mm without cyclic effect
2 mm with/out cyclic effect
0.12 5 mm with cyclic effect
5 mm without cyclic effect
Friction coefficient [-]

0.1

0.08

0.06

0.04

0.02
0 20 40 60 80 100 120 140 160 180 200
Velocity [mm/s]

Figure 5.20: Scatter from the Laboratory test of the bearing for 1 Mpa

77
0.14
1 mm with cyclic effect
1 mm without cyclic effect
0.12
2 mm with cyclic effect
2 mm without cyclic effect
0.1 5 mm with cyclic effect
Friction coefficient [-]

5 mm without cyclic effect

0.08

0.06

0.04

0.02

0
0 20 40 60 80 100 120 140 160 180 200
Velocity [mm/s]

Figure 5.21: Scatter from the Laboratory test of the bearing for 5 Mpa

0.14
1 mm with cyclic effect
1 mm without cyclic effect
2 mm with cyclic effect
0.12 2 mm without cyclic effect
1 mm with cyclic effect
5 mm without cyclic effect
0.1
Friction coefficient [-]

0.08

0.06

0.04

0.02

0
0 20 40 60 80 100 120 140 160 180 200
Velocity [mm/s]

Figure 5.22: Scatter from the Laboratory test of the bearing for 10 Mpa

0.14
1 mm, 1 MPa
1 mm, 5 MPa
0.12 1 mm, 10 MPa
5 mm, 1 MPa
5 mm, 5 MPa
0.1
5 mm, 10 MPa
Friction coefficient [-]

0.08

0.06

0.04

0.02

0
0 20 40 60 80 100 120 140 160 180 200
Velocity [mm/s]

Figure 5.23: Results from both experiments 5 mm and 1 mm for the different pressures

78
5.7 Conclusion

The following conclusions have been reached as a result of what have been explained whitin
the previous chapters:

• The proposed modification of the classical Bouc-Wen model produces an improvement in


terms of model’s parameter definition

• The proposed modification of the Bouc-Wen model facilitates the generalization of the
model when it is used to simulate the dissipation of energy due to friction in pot bearings

• 3D models produce almost the same response as the SDOF models, suggesting that the
dissipation of energy at pot bearings is mainly determined by the interaction PTFE-steel
and that the rest of interactions could be even neglected

• The agreement in the PTFE-steel contact 3D models differs for different velocities and
degrees of lubrication. Small velocities have high concordance while deviations up to 50%
are observed for high velocities.

5.8 Further questions and considerations

The following issues are still unanswered and they will probably be the key to an accurate
solution for the problems exposed in this thesis. They are organized in a priority order in the
opinion of the authors:

• Thermomechanical properties of the PTFE.

• Mechanical properties of the PTFE in the tangential direction.

• Micro-mechanics of the PTFE. This thesis has been carried out in a ”macro-scale” context,
is there some issues in the ”micro-sale” that cannot be neglected?

• How sensitive is the interaction PTFE-steel to the type and amount of lubricant?

• What are the scale factors between pot-bearings of different size? Can the current models
be used in larger bearings? This can be verified in two ways; directly by testing larger
bearings and/or indirectly by testing existing bridges.

• Increasing accuracy of current 3D models. During the realization of this thesis some
simplifications have been carried out in order to save computational time in the 3D
models. Increasing the detail in the models could bring new information about the topic.
In addition, increasing the accuracy of the model could unequivocally answer if the PTFE-
steel is strongly the main factor to consider when analyzing the dissipation of energy due
to friction at pot bearing or if there are other sources.

• An implementation in real bridges would be the final stage of this study. It could be
done in several ways, such as using a user element subroutine that gathers all of this
information or implementing friction models with subroutines.

79
Bibliography

[1] Anil K. Chopra. Dynamic of Structures. Theory and Applications to Earthquake Engi-
neering. University of California at Berkeley. Prentice Hall, 1: 65-125, 2011.

[2] M. Ülker-Kaustell and R. Karoumi. Application of the continuous wavelet transform on


the free vibrations of a steel-concrete composite railway bridge. Engineering Structures,
33:911–919, 2011.

[3] M. Ülker-Kaustell and R. Karoumi. Influence of non-linear stiffness and damping on


the train-bridge resonance of a simply supported railway bridge. Engineering Structures,
41:350–355, 2012.

[4] ERRI D 214 RP3, Rail Bridges for Speeds > 200 km/h - Recomendations for calculating
damping in rail bridge decks. European Rail Research Institute, 2000.

[5] CEN. Eurocode 1: Actions on structures - Part 2: Traffic loads on bridges. CEN, Euro-
pean Comittee for Standardization, 2008.

[6] M.Ülker-Kaustell, R. Karoumi, and C. Pacoste. Simplified analysis of the dynamic


soil-structure interaction of a portal frame railway bridge. Engineering Structures,
32:3692–3698, 2010.

[7] M. Ülker-Kaustell. Some aspects of the dynamic soil-structure interaction of a portal


frame railway bridge. TRITA-BKN. Licentiate thesis, 102, 2009.

[8] Johan Östlund. Soil-structure interaction of end-frames for high-speed railway bridges.
Degree project in Civil engineering and Urban management in Kungliga Tekniska
Högskolan, 2016.

[9] Influence of rate-independent hysteresis on the dynamic response of a railway bridge.


International Journal of Rail Transportation. 1(4) 237-257, 2013.

[10] M. Dolce, D. Cardone, and F. Croatto. Frictional behavior of steel-ptfe interfaces for
seismic isolation. Bulletin of Earthquake Engineering, 3:75–99, 2005.

[11] Swedish Standards Institute (SIS). SS-EN 1337-2:2004 Part 2: Sliding elements.

[12] Swedish Standards Institute (SIS). SS-EN 1337-2:2004 Part 5: Pot bearings.

[13] Dassault Systémes. Abaqus Analysis 6.11 User’s Manual. 2011

[14] Coulomb C.A. The theory of simple machines (in french). Mem. Math. Phys. Acad. Sci.,
10:161–331, 1785.

80
[15] Euler L. Sur la diminution de la resistance du frottement. Mem. Acad. Sci. Berlin,
4:133–148, 1748a.

[16] Mostaghel, N. and Tanbakuchi, J.T. Response of Sliding Structures to Earthquake Sup-
port Motion. Earthquake Engineering and Structural Dynamics. 11, 729–748, 1983.

[17] Fan, F.G., Ahmadi, G. and Tadjbakhsh, I.G. 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, 1988.

[18] Constantinou, M.C., Caccese, J. and Harris, H.G. Friction characteristics of PTFE– steel
interfaces under dynamic conditions. Earthquake Engineering and Structural Dynamics.
15(6), 751–759, 1987.

[19] Hwang, J.S., Chang, K.C. and Lee, G.C. Quasi-static and dynamic characteristics of
PTFE-stainless interfaces. Journal of Structural Engineering 116(10), 2747–2762, 1990.

[20] Mokha, A., Constantinou, M. and Reinhorn, A.M. PTFE bearings in base isolation:
testing. Journal of Earthquake Engineering 116(2), 438–454, 1990.

[21] Tyler, R.G. Dynamic tests on PTFE sliding layers under earthquake conditions. Bulletin
of the New Zealand National Society for Earthquake Engineering 10(3), 129–138, 1977.

[22] Wen, Y.K. Method for Random Vibration of Hysteretic Systems. Journal of the Engi-
neering Mechanic Division, ASCE, 102 (EM2), 1976.

[23] Chang J.C., Hwang J.S. and Lee G.C. Analytical model for sliding behaviour of Teflon-
stainless steel interfaces. Journal of Engineering Mechanics 116, 2749–2763, 1990.

[24] Constantinou, M., Mokha, A. and Reinhorn, A.M. PTFE bearings in base isolation:
modelling. Journal of Earthquake Engineering 116(2), 455–472, 1990.

[25] E. Rabinowicz. The Nature of the Static and Kinematic Coefficients of Friction. Lubri-
cation Laboratory, Massachusetts Institute of Technology, Cambridge, 22(11) 1951.

[26] V.I. Johannes, M.A. Green, and C.A. Brockley. The role of the rate of application of the
tangential force in determining the static force coefficient. Wear, 24:381-385, 1973

[27] S. Bjorklund. A random model for micro-slip between nominallyflat surfaces. Trans.
ASME, J. of Tribology, 119(4):726–732, 1997.

[28] E. Rabinowicz. Stick and Slip. Scientific American, 1956.

[29] G. Loamiento, N.Bonessio & G. Benzoni. Friction Model for Sliding Bearing under Seismic
Excitation. Journal of Earthquake Engineering, 17:8, 1162-1191, 2013.

[30] Peter Wriggers. Computational Contact Mechanics. University of Hannover, 2:11-18,


2002.

[31] F.E. Kennedy, X. Tian, The effect of interfacial temperature on friction and wear of
thermoplastics in the thermal control regime, Tribology Series 27 (C), pp. 235–244. 1994.

[32] I. Tzanakis, N. Garland, M. Hadfield, Cavitation damage incubation with typical fluids
applied to a scroll expander system, Tribology International 44 (12), pp.1668–1678. 2011.

81
[33] F.J. Martı́nez, M. Canales, J.M. Bielsa, M.A. Jiménez, Relationship between wear rate
and mechanical fatigue in sliding TPU-metal contacts, Wear 268 (2–3) , pp. 388–398.
2010.

[34] W. Wieleba, The role of internal friction in the process of energy dissipation during PTFE
composite sliding against steel, Wear 258 (5–6), pp. 870–876. 2005.

[35] B.N.J. Persson, Rubber friction: role of the flash temperature, Journal of Physics: Con-
densed Matter 18 (32), pp. 7789–7823. 2006.

[36] I. Tzanakis ,M. Conte , M. Hadfield , T.A. Stolarski, Experimental and analytical thermal
study of PTFE composite sliding against high carbon steel as a function of the surface
roughness, sliding velocity and applied load Wear 303, pp. 154–168. 2003.

[37] F.P. Bowden, D. Tabor, The Friction and Lubrication of Solids, Clarendon Press, Oxford,
1964.

[38] Conte M., B. Fernandez, A. Igartua, Effect of surface temperature on tribological behavior
of PTFE composites, in: Proceedings of the Surface Effects and Contact Mechanics X,
WIT press, UK, pp. 219–230. 2011.

[39] S. Bahadur, The development of transfer layers and their role in polymer tribology, Wear
245, pp. 92–99. 2000.

[40] J. Khedkar, I. Negulescu, E.I. Meletis, Sliding wear behavior of PTFE composites, Wear
252, pp. 361–369. 2002.

[41] B. D. Coleman and W. Noll. The thermodynamics of elastic materials with heat conduc-
tion and viscosity. Archive for Rational Mechanics and Analysis 13, 167-178, 1963.

[42] Gerhard A. Holzapfel. Nonlinear Solid Mechanics - A continuum approach for engineering.
1st edition. Wiley. 4, 205-304, 2000.

[43] C. Truesdell and W. Noll. The non-linear field theories of mechanics. 2nd edition. Springer-
Verlag, Berlin. 1-6, 1992.

[44] R.W. Ogden. Elastic deformations of rubberlike solids, in: H.G. Hopkins, and M.l. Sewell,
cds., Mechanics of Solids, the Rodney Hill 60th Anniversary Volume. Pergamon Press.
Oxford. (6) 499-537, 1982.

[45] R.W. Ogden. Recent advances in the phenomenological theory of rubber elasticity, Rubber
Chemistry and Technology. 59 (6), 261-383. 1986.

[46] Jean Lemaitre. Handbook of Materials Behaviour Models . 1st edition. 2:2, 69-95, 2001.

[47] Peter Wriggers. Nonlinear Finite Element Methods. 1st edition. Springer, Germany. 2008.

[48] Hertz H. Über die Berührung fester elastischer Körper. Journal für die Reine und Ange-
wandte Mathematik, 29:156–171, 1882.

[49] M. Mooney. A theory of large elastic deformation, JOurnal of Applied Physisc 11, 58 593
(6), 1940.

[50] R.S. Rivlin. Large elastic deformations of isotropics materials, V. The problem of flexure,
Proceedings of the Royal Society of London A195, 463-473, (6), 1949.

82
[51] L. Anand A constitutive model for compressible elastomeric solids, Computational Me-
chanics 18, 339-355, (6), 1996.

[52] I. Caliò, F. Cannizzaro, B. Pantò. A macro-element approach for modeling the nonlin-
ear behaviour of monumental buildings under static and seismic loadings. University of
Catania, Italy. 15th World Conference on Earthquake engineering. Lisbon, 2012.

[53] R. Bouc. Modèle mathématique d’hystérésis.Acustica, 24:16–25, 1971.

[54] K. C. Valanis. A theory of viscoplasticity without a yield surface. part i –general theory.
Archives of Mechanics, 23:517–533, 1971.

[55] K. Ishihara.Soil behaviour in earthquake geotechnics. Oxford University Press,London,


1996.

[56] F. Ikhouane and José Rodellar. Dynamic properties of the hysteretic Bouc-Wen model,
Systems & Control Letters, 2006.

[57] Erlicher, S., Point, N., “Thermodynamic admissibility of Bouc-Wen type hysteresis mod-
els,” C. R. Mecanique 332, pp. 51-57, 2004.

[58] Ma, F., Zhang, H., Bockstedte, A., Foliente, G.C. and Paevere, P. “Parameter Analysis
of the Differential Model of Hysteresis,” ASME Journal of Applied Mechanics 71, pp.
342-349, 2004.

[59] Aristotelis E. Charalampakis. Parameters of Bouc-Wen Hysteretic model Revisited.


School of Civil Engineering National Technical University of Athens. 9th HSTAM In-
ternational Congress on Mechanics Limassol, Cyprus, July, 2010.

[60] Constantinou, M.C., Adnane, M.A., “Dynamics of soil-base-isolated structure systems:


evaluation of two models for yielding systems,” Report to NSAF, Department of civil
engineering, Drexel University, Philadelphia, 1987.

[61] Charalampakis, A.E., Koumousis, V.K. “Identification of Bouc-Wen hysteretic systems


by a hybrid evolutionary algorithm,” J. of Sound and Vibration 314, pp.571-585, 2008.

[62] P.J. Raea, D.M. Dattelbaumb. The properties of poly(tetrafluoroethylene) (PTFE) in


compression. Los Alamos National Laboratory, Structure Property Relations, MST-8,
MS-G755, Los Alamos, NM 85745, USA, 2004.

[63] Properties Handbook of PTFE DU Pont, 2014.

[64] G. M. Calvi, K. Kawashima, I. Billings, A. Elnashai, C. Nuti, A. Pecker, P. E. Pinto, N.


M. J. Priestley, M. Rodriguez, L. Di Sarno, P. Franchin, D. Pietra, I. Vanzi, fib Bulletin
39, Seismic bridge design and retrofit – structural solutions, 6.3.3. Sliding devices, pp.
86-87. 2007

83

You might also like