Вы находитесь на странице: 1из 228

Hydraulic Engine Mount Modeling,

Parameter Identification and


Experimental Validation

by

Aaron A. Geisberger

A thesis
presented to the University of Waterloo
in fulfillment of the
thesis requirement for the degree of
Master of Applied Science
in
Mechanical Engineering

Waterloo, Ontario, 2000


c Aaron A. Geisberger, 2000

I hereby declare that I am the sole author of this thesis.


I authorize the University of Waterloo to lend this thesis to other institutions or individuals for
the purpose of scholarly research.

Aaron A. Geisberger

I authorize the University of Waterloo to reproduce this thesis by photocopying or other means,
in total or in part, at the request of other institutions or individuals for the purpose of scholarly
research.

Aaron A. Geisberger

ii

The University of Waterloo requires the signatures of all persons using or photocopying this
thesis. Please sign below, and give address and date.

iii

Abstract
This thesis presents a modeling study of hydraulic engine mounts currently used in the
automotive industry. Nonlinear model aspects are developed and used with experimentally identified parameters to validate the model response characteristics. It is felt that this contribution
will help engineers in reducing mount design time, by providing insight into the effects of various
parameters within the mount.
In the thesis, various components and flow passages are assigned lumped parameters, in
order to arrive at a physical representation of the mount system. Bond graph techniques are used
to derive the differential system equations, which are used to develop the mechanical analogue
of the system. Of particular interest are the dynamic behaviors of the decoupler, inertia track
and bell assemblies. The nonlinear responses of these systems are considered, and in some cases
produce interesting behaviors that are typical of nonlinear systems.
An extensive set of experiments are conducted using a unique test apparatus to provide
numerical data for the various parameter identification techniques. These data also establish the
relative importance of several damping, inertia and stiffness terms. In addition, the measured
responses of the mounts to loading at various frequencies and amplitudes are compared to the
predictions of the mathematical model. The comparisons generally show a very good agreement
(better than 10%), which indicate that the analysis has been effective.
It is shown that the physical modeling in this thesis provides the appropriate system response over the full range of loading conditions (frequency and amplitude) encountered in practice. These results present an improvement on existing nonlinear models, which were limited
either to low frequency, high amplitude conditions, or use a piecewise linear approach. The results show the importance of several individual design parameters on the performance of the
mounts. Some areas requiring additional work, such as the fluid resistance terms for the inertia
track, and column inertia data for the decoupler orifices, are identified.

iv

Acknowledgments
I would like to thank my supervisors Dr. M.F. Golnaraghi and Dr. A. Khajepour, for their
input and support over the duration of this project. Also, the effort from my readers, Dr. G. Heppler and Dr. R. Macdonald, to promptly read my thesis was greatly appreciated.
Thanks are due for the financial support from Cooper-Standard Automotive, the Ontario
Graduate Scholarships and the Natural Sciences and Engineering Research Council (NSERC).
The cooperation and enthusiasm of everyone in the NVH Control Systems group at CooperStandard Automotive has made this research possible. Particularly, I would like to thank Rob
Bender, Carl Ohrling, Rob Sehn, Bernie Rice and Rob Paladichuk for all their help. Enjoy the
read.
I would like to take this opportunity to thank my parents, for their comfort and support
through all my endeavors, and the values they have taught me. Thanks also to my sister and
brother for providing happiness and inspiration to my life.
This thesis would not have been possible without the encouragement of friends, and the
many festivities. To: Barry & Tara, Brent & Jen, Dag, Gord, Gus, Shelley, Taylor, Tom & Kelly
and Thanh, I am very fortunate to have you as friends, cheers. Also, to my roommates, Kyle,
Matt, Mike and Trev, you kept life interesting during this work, thanks.
Now, I am off to windsurf.

Table of Contents
1 Introduction

1.1 Desired Engine Mount Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


1.1.1 Road Excitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.2 Engine Eccentricity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 The Hydraulic Engine Mount . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.1 Determining Mount Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.2 Mount Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.4 Objectives and Thesis Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2 Mathematical Modeling of Hydraulic Engine Mounts

17

2.1 Linear Model of a Typical Hydraulic Mount . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17


2.1.1 Parameter and Variable Assignment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.1.2 Bond Graph Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.1.3 System Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.1.4 Developing Two Linear Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.2 Mechanical System Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3 Nonlinear Model Extension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.3.1 Upper Chamber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

vi

2.3.2 Lower Chamber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30


2.3.3 Inertia Track . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.3.4 Decoupler . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.4 Modeling the Additional Bell System. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
2.4.1 Lumped Parameter Assignment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
2.4.2 Bond Graph Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
2.4.3 System Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
2.5 Mechanical System Model with a Bell . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
2.6 Nonlinear Enhancement of the Bell Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
2.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

3 An Insight into the Mathematical Models

57

3.1 Linear Model Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57


3.2 Parameter Study for Linear Model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.2.1 Compliance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.2.2 Inertia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.2.3 Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
3.2.4 Effective Pumping Area. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.2.5 Linear Stiffness and Damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
3.3 Nonlinear Decoupler Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
3.3.1 Low Frequency Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
3.3.2 High Frequency Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
3.3.3 Influence of Nonlinear Decoupler Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
3.4 Linear Bell Model Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.5 Parameter Studies on a Mount with the Bell System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
3.5.1 Compliance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
vii

3.5.2 Inertia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.5.3 Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
3.5.4 Area Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
3.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

4 Experimental Design

93

4.1 Goals and Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93


4.2 Design Criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.3 Final Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
4.4 Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4.5 Calibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
4.5.1 Vessel Chamber Compliance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
4.6 Apparatus Capabilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

5 Parameter Identification

105

5.1 Upper Chamber Compliance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105


5.1.1 Stiffness and Damping Parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
5.1.2 Effective Pumping Area Parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.1.3 Volumetric Compliance Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.2 Lower Chamber Compliance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.3 Inertia Track Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
5.4 Decoupler Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
5.5 Bell Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
5.5.1 Fluid Column Inertia and Resistance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
5.5.2 Chamber Compliance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
5.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

viii

6 Simulation and Validation of System Models

136

6.1 Simulation Techniques. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136


6.2 Model Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
6.2.1 Inertia Track . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
6.2.2 Decoupler . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
6.2.3 Inertia Track and Decoupler . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
6.2.4 Inertia Track and Bell . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
6.2.5 Inertia Track, Decoupler and Bell . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
6.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

7 Preliminary Studies on Parameter to Geometry Relation

153

7.1 Inertia Track . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153


7.1.1 Inertia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
7.1.2 Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
7.2 Decoupler. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
7.2.1 Inertia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
7.2.2 Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
7.2.3 Switching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
7.3 Bell Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
7.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163

8 Conclusions and Recommendations


8.1

164

Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

8.2 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169

ix

A Bond Graph Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172


B Frequency Domain Equation Development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
C Detailed Drawings of Experimental Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
D The Extended Kalman Filter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203

List of Tables
Table 3.1:

Linear model parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

Table 3.2:

Nonlinear model parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

Table 3.3:

Parameters used in linear bell model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

Table 3.4:

Frequency and mode shape vectors of high frequency bell model. . . . . . . . . . . . . . . 79

Table 3.5:

Frequency modifications to changes in the upper chamber compliance


parameter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

Table 3.6:

Frequency modifications to changes in the bell chamber compliance parameter. . 84

Table 3.7:

Frequency modifications resulting from changes in the decoupler fluid column


inertia parameter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

Table 3.8:

Frequency modifications resulting from changes in the bell fluid column inertia
parameter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

Table 3.9:

Frequency modifications resulting from changes in the mount area parameter. . . 89

Table 3.10: Frequency modifications resulting from changes in the effective pumping area
parameter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Table 3.11: Frequency modifications resulting from changes in bell area parameter. . . . . . . . . 90
Table 5.1:

Test conditions used for dynamic characterization, at 1700 N preload. . . . . . . . . . 107

Table 5.2:

Inertia track, decoupler and bell parameters identified for the Cooper-Standard
mount. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

Table 7.1:

Calculated versus measured inertia parameters for the inertia track. . . . . . . . . . . . 154

Table 7.2:

Measured resistance parameters of the inertia track. . . . . . . . . . . . . . . . . . . . . . . . . . . 155

Table 7.3:

Effective inertia and fluid column height for a simple orifice. . . . . . . . . . . . . . . . . . 156

Table 7.4:

Calculated versus measured inertia for the complete decoupler system. . . . . . . . . 158

Table 7.5:

Measured decoupler resistance parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159

Table 7.6:

Calculated decoupler switching parameters versus measured. . . . . . . . . . . . . . . . . . 161

Table 7.7:

Measured bell parameter, indicating fluid column height. . . . . . . . . . . . . . . . . . . . . . 162

xi

List of Tables
Table 1.1:

Bondgraph Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174

Table 1.2:

Causality of Bondgraph Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175

xii

List of Figures
Figure 1.1:

Low frequency excitations: (a) SDOF base excitation representation; (b)


Frequency domain relative displacement transmissibility. . . . . . . . . . . . . . . . . . . . . . . 3

Figure 1.2:

High frequency excitations: (a) SDOF forced input excitation; (b) transmitted
force frequency domain plot. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

Figure 1.3:

Cross section of a typical hydraulic engine mount. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

Figure 1.4:

Standard test configuration for hydraulic engine mounts. . . . . . . . . . . . . . . . . . . . . . . 7

Figure 1.5:

Illustration of MTS 1000-Hz rate machine used to measure dynamic


characteristics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

Figure 1.6:

Transmitted force and excitation vectors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

Figure 1.7:

Typical mount dynamic characteristics: (a) Low frequency, large amplitude; (b)
High frequency, small amplitude. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

Figure 2.1:

Cross section of typical hydraulic mount. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

Figure 2.2:

Lumped parameter system model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

Figure 2.3:

Bond graph model for a typical hydraulic engine mount. . . . . . . . . . . . . . . . . . . . . . 20

Figure 2.4:

Transmitted force to the mount base: (a) decoupler closed; (b) free floating
decoupler. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

Figure 2.5:

Mechanical representation of the high frequency, low amplitude linear model. . 25

Figure 2.6:

Transmitted force to the base in the mechanical model. . . . . . . . . . . . . . . . . . . . . . . . 25

Figure 2.7:

Mechanical model that includes the decoupler and inertia track.. . . . . . . . . . . . . . . 27

Figure 2.8:

Capturing bulge damping in the upper chamber: (a) previous model; (b) addition
of a resistance term in the compliance.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

Figure 2.9:

Bond graph model of the compliance system with resistive parameter.. . . . . . . . . 30

Figure 2.10: The nonlinear decoupler resistance: (a) sign convention; (b) switching
function. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

xiii

List of Figures
Figure 2.11: Additional resistance function indicating influence of parameter. . . . . . . . . . . . . . . 36
Figure 2.12: Illustration of the arctangent function. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
Figure 2.13: Isolated decoupler model that includes cage geometry: (a) geometry parameters;
and (b) pressure gradient breakdown. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
Figure 2.14: The decoupler area: (a) sign convention; (b) switching function. . . . . . . . . . . . . . . 40
Figure 2.15: Decoupler leak flow: (a) system model and sign convention; (b) nonlinear
resistance function. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
Figure 2.16: Cross section of a hydraulic mount with a bell system. . . . . . . . . . . . . . . . . . . . . . . . 44
Figure 2.17: Lumped parameter fluid system model including the bell. . . . . . . . . . . . . . . . . . . . . 45
Figure 2.18: Bond graph model that includes the inertia track, decoupler and bell. . . . . . . . . . . 46
Figure 2.19: Mechanical lever at top of mount. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
Figure 2.20: Mechanical model of the hydraulic mount with a bell. . . . . . . . . . . . . . . . . . . . . . . . . 50
Figure 2.21: Mechanical model indicating transmitted force points.. . . . . . . . . . . . . . . . . . . . . . . . 51
Figure 3.1:

Dynamic stiffness and phase response of the linear low frequency model. . . . . . 59

Figure 3.2:

Low frequency linear model response of: (a) volume passing through inertia
track; and (b) pressure in upper chamber. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

Figure 3.3:

Low frequency linear model response, showing the stiffness response and liquid
column motion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

Figure 3.4:

High frequency linear model response, showing the stiffness response and liquid
column motion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

Figure 3.5:

Linear model response to changes in the upper chamber compliance parameter. 64

Figure 3.6:

Linear model response to changes in the fluid inertia parameter. . . . . . . . . . . . . . . 65

Figure 3.7:

Linear model response to changes in the fluid resistance parameter. . . . . . . . . . . . 66

Figure 3.8:

Linear model response for changes in the effective pumping area parameter. . . . 67

Figure 3.9:

Linear model response to changes in the linear stiffness parameter. . . . . . . . . . . . . 68

Figure 3.10: Time domain simulation of nonlinear decoupler model: (a) decoupler volume;
(b) inertia track volume; (c) upper chamber pressure. . . . . . . . . . . . . . . . . . . . . . . . . . 71
Figure 3.11: Low frequency simulation of nonlinear model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
Figure 3.12: Dynamic stiffness characteristics of the nonlinear model under high frequency
excitations.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
Figure 3.13: Decoupler motion in the nonlinear model under high frequency excitations. . . . 73
Figure 3.14: Decoupler motion indicating region on instability and dependence on sine sweep
conditions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
Figure 3.15: Dynamic stiffness and phase indicating impact of sine sweep conditions. . . . . . . 75
xiv

List of Figures
Figure 3.16: Simulation of decoupler volume indicating the response to changes in the
decoupler parameter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
Figure 3.17: Dynamic stiffness and phase characteristics from the simulated linear model that
includes the bell and decoupler. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
Figure 3.18: High frequency linear model response of: (a) decoupler fluid column motion;
and (b) bell fluid column motion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
Figure 3.19: High frequency linear model response of: (a) upper chamber pressure; and (b)
bell chamber pressure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
Figure 3.20: Dynamic stiffness and phase response to changes in the upper chamber
compliance parameter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Figure 3.21: Dynamic stiffness and phase response to changes in the bell chamber compliance
parameter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Figure 3.22: Dynamic stiffness and phase response to changes in the decoupler fluid column
inertia parameter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
Figure 3.23: Dynamic stiffness and phase response to changes in the bell fluid column inertia
parameter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
Figure 3.24: Dynamic stiffness and phase response to changes in the decoupler resistance
parameter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
Figure 3.25: Dynamic stiffness and phase response to changes in the mount area parameter. 89
Figure 3.26: Dynamic stiffness and phase response to changes in the effective pumping
area. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
Figure 3.27: Dynamic stiffness and phase response to changes in the bell area parameter. . . . 91
Figure 4.1:

Experimental Approach. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

Figure 4.2:

Schematic of experimental test apparatus. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

Figure 4.3:

Photo of experimental apparatus in MTS servo-controlled rate machine. . . . . . . . 97

Figure 4.4:

Photograph of experimental apparatus. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

Figure 4.5:

Instrumentation setup for the experimental apparatus. . . . . . . . . . . . . . . . . . . . . . . . . 99

Figure 4.6:

Illustration of pressure sensor calibration tool. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

Figure 4.7:

Illustration of the volumetric compliance associated with the vessel lower


chamber. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

Figure 4.8:

Plot of lower chamber pressure versus actuator displacement. . . . . . . . . . . . . . . . . 102

Figure 5.1:

Cross section of the Cooper-Standard mount under investigation. . . . . . . . . . . . . 106

Figure 5.2:

Static force versus displacement. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

xv

List of Figures
Figure 5.3:

Stiffness and damping properties measured at various excitation conditions: (a)


low frequency high amplitude; (b) high frequency low amplitude. . . . . . . . . . . . . 108

Figure 5.4:

Test configuration for measuring the effective area and volumetric compliance
parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

Figure 5.5:

Plot of effective pumping area versus mount displacement. . . . . . . . . . . . . . . . . . . 109

Figure 5.6:

Static curve of volume versus pressure in the upper chamber. . . . . . . . . . . . . . . . . 111

Figure 5.7:

Upper compliance parameters identified at 15 psi mean pressure using the test
apparatus: (a) volumetric compliance; and (b) bulge damping. . . . . . . . . . . . . . . . 113

Figure 5.8:

Dynamic stiffness and phase response of the special mount, measured using
the test conditions in Table 5.1: (a) low frequency; and (b) high frequency
response.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

Figure 5.9:

Upper compliance parameters identified using the system model and test
conditions in Table 5.1: (a) volumetric compliance; and (b) bulge damping. . . 114

Figure 5.10: Test configuration for determining the volumetric compliance of the lower
chamber. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
Figure 5.11: Static measurements of lower chamber volumetric expansion versus pressure. 115
Figure 5.12: Test configuration for isolating the inertia track paramters. . . . . . . . . . . . . . . . . . . . 117
Figure 5.13: Parameter trends with respect to inertia track flow, for: (a) inertia; and (b)
resistance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
Figure 5.14: Isolated inertia track excitation plots including: (a) time domain segment
of random flow input; and (b) power spectral density of complete flow
perturbation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
Figure 5.15: Time domain segment of measured pressure differential versus model output
using estimated parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
Figure 5.16: Test configuration for isolating the decoupler parameters.. . . . . . . . . . . . . . . . . . . . 122
Figure 5.17: Steady-state time domain data of actual decoupler volume and pressure
differential at 12 Hz, showing: (a) nonlinear cage contact; and (b) linear
uncoupled response. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
Figure 5.18: Parameter trends with respect to decoupler flow, for: (a) fluid column inertia;
and (b) resistance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
Figure 5.19: Time segment of the decoupler input flow and pressure differential, indicating
locations where leak resistance is identified. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
Figure 5.20: Decoupler state and parameter identification: (a) pressure differential across
decoupler; (b) decoupler volume state variable; and (c) decoupler switching
parameter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

xvi

List of Figures
Figure 5.21: Close up of the decoupler state and parameter identification: (a) pressure
differential across decoupler; (b) decoupler volume state variable; and (c)
decoupler switching parameter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
Figure 5.22: Cross section of Cooper-Standard mount that includes a bell plate. . . . . . . . . . . . 131
Figure 5.23: Test configuration for isolating the bell parameters. . . . . . . . . . . . . . . . . . . . . . . . . . 131
Figure 5.24: Parameter trends with respect to bell flow, for: (a) inertia; and (b) resistance. . 133
Figure 5.25: Illustration of the quarter plate model used to approximate the chamber
compliance, showing: (a) boundary constraints and 3 MPa evenly distributed
pressure load; (b) plate deflection in linear elastic region. . . . . . . . . . . . . . . . . . . . . 134
Figure 6.1:

Actuator model used to provide excitations with zero initial conditions. . . . . . . 138

Figure 6.2:

Illustration of the Cooper-Standard mount being compared with simulation


data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

Figure 6.3:

Low frequency simulation versus measured response, on a mount with only an


inertia track. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

Figure 6.4:

High frequency simulation versus measured response, on a mount with only a


decoupler. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

Figure 6.5:

Low frequency simulation versus measured response, on a mount including a


decoupler and inertia track. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145

Figure 6.6:

High frequency simulation versus measured response, on a mount including a


decoupler and inertia track. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145

Figure 6.7:

Illustration of the Cooper-Standard mount used to validate system models with a


bell. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147

Figure 6.8:

Low frequency simulation versus measured response, on a mount including an


inertia track and bell system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147

Figure 6.9:

High frequency simulation versus measured response, on a mount including an


inertia track and bell system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

Figure 6.10: Low frequency simulation versus measured response, on a mount including an
inertia track, decoupler and bell system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
Figure 6.11: High frequency simulation versus measured response, on a mount including an
inertia track, decoupler and bell system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
Figure 6.12: Comparison of high frequency response with and without the bell plate, using
both measured and modeled response. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
Figure 7.1:

Illustration of inertia track geometries under investigation. . . . . . . . . . . . . . . . . . . . 154

Figure 7.2:

Measured resistance parameters for each test case: (a) flow resistance; and (b)
velocity damping. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

xvii

List of Figures
Figure 7.3:

Fluid inertia tests of the decoupler orifice with: (a) seven holes; and (b) nineteen
holes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

Figure 7.4:

Illustration of the decoupler inertia with a decoupler plate. . . . . . . . . . . . . . . . . . . . 158

Figure 7.5:

Measured decoupler resistance parameters plotted for each of the decoupler


cases investigated. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

Figure 7.6:

Measured decoupler displacement versus the decoupler switching parameter. . 161

Figure 7.7:

Illustration of decoupler geometry used to calculate the switching parameters. 162

xviii

Nomenclature
Ab

annular bell area [mm2 ]

Ad

decoupler area [mm2 ]

Ad_holes

area of the decoupler holes [mm2 ]

Ad_f nc

function for nonlinear decoupler area [mm2 ]

Ai

cross-sectional area of the inertia track [mm2 ]

Am

area of mount at the bell [mm2 ]

Ap

effective pumping area of the upper compliance [mm2 ]

Apiston

piston area on the test apparatus [mm2 ]

bb

damping coefficient of the bell [kg/s]

bd

damping coefficient of the decoupler [kg/s]

bi , bi1

damping coefficient of the inertia track [kg/s]

bi2

damping parameter on squared velocity of the inertia track [kg/mm]

br

damping coefficient of the upper compliance [kg/s]

Cb

compliance of the bell chamber [mm5 /N]

Crig

lower chamber compliance of test apparatus [mm5 /N]

C1

compliance of the upper chamber [mm5 /N]

xix

Nomenclature
C2

compliance of the lower chamber [mm5 /N]

damping parameter in the SDOF system [kg/s]

Fp

preload force [N]

FT

transmitted force to mount base [N]

FT

complex amplitude of the transmitted force to mount base [N]

fn

natural frequency [Hz]

fo

forcing coefficient [mm2 /s2 ]

hf luid

approximate fluid column height [mm]

hgap

gap height within the decoupler cage [mm]

Ib

fluid inertia of the bell [kg/mm4 ]

Id

fluid inertia of the decoupler [kg/mm4 ]

Ii

fluid inertia of the inertia track [kg/mm4 ]

cross point dynamic stiffness [N/mm]

stiffness parameter in the SDOF system [N/mm]

kb

elastic stiffness of the bell chamber [N/mm]

kr

elastic stiffness of the upper compliance [N/mm]

k1

elastic stiffness of the upper chamber [N/mm]

k2

elastic stiffness of the lower chamber [N/mm]

length of the inertia track [mm]

Mplate

mass of the decoupler plate [kg]

mass parameter in the SDOF system [kg]

mb

mechanical mass of the bell inertia [kg]

md

mechanical mass of the decoupler inertia [kg]

mi

mechanical mass of the inertia track [kg]


xx

Nomenclature
Pb

pressure in the bell chamber [N/mm2 ]

Ps1

lower chamber pressure sensor reading on the test apparatus [N/mm2 ]

Ps2

upper chamber pressure sensor reading on the test apparatus [N/mm2 ]

P1

pressure in the upper chamber [N/mm2 ]

P2

pressure in the lower chamber [N/mm2 ]

Qactual

calculated flow through the holding plate in the test apparatus [mm3 /s]

Qb

flow through the bell [mm3 /s]

b
Q

complex amplitude of flow through the bell [mm3 /s]

Qd

flow through the decoupler [mm3 /s]

d
Q

complex amplitude of flow through the decoupler [mm3 /s]

Qexcit

excitation flow in test apparatus [mm3 /s]

Qi

flow through the inertia track [mm3 /s]

QL

leak flow through the decoupler [mm3 /s]

QT

flow into the compliant upper chamber [mm3 /s]

Radd

additional flow stopping resistance [kg/s-mm4 ]

Rb , Rb1

fluid flow resistance across the bell [kg/s-mm4 ]

Rb2

resistance parameter on squared flow across the bell [kg/mm7 ]

Rcb

flow resistance for bell chamber bulge damping [kg/s-mm4 ]

Rd , Rd1

fluid resistance parameter on squared flow across the decoupler [kg/s-mm4 ]

Rd2

resistance parameter on squared flow across the decoupler [kg/mm7 ]

Rd3

decoupler flow stopping parameter [mm1 ]

Ri , Ri1

resistance across the inertia track [kg/s-mm4 ]

xxi

Nomenclature
Ri2

resistance parameter on squared flow across the inertia track [kg/mm7 ]

RL2

leak flow resistance [kg/mm7 ]

RL2_f nc

nonlinear function for leak flow resistance [kg/mm7 ]

R1

flow resistance for upper chamber bulge damping [kg/s-mm4 ]

time step for discrete data [s]

matrix of inputs for least-squares method

Vactual

calculated volume of fluid though the holding plate in test apparatus [mm3 ]

Vactual

complex amplitude of volume though the holding plate [mm3 ]

Vd

volume of fluid through the decoupler [mm3 ]

Vd_ max

maximum volume of fluid accommodated through the decoupler [mm3 ]

Vexcit

excitation volume in the test apparatus [mm3 ]

Vi

velocity of the inertia track [mm/s]

VT

volume of displaced into the upper chamber [mm3 ]

VT

complex amplitude of volume displaced into the upper chamber [mm3 ]

wdr

driving frequency [rad/s]

wn

undamped natural frequency [rad/s]

amplitude of excitation displacement [mm]

amplitude of vibration of the excitation [mm]

Xact

actuator displacement on test apparatus [mm]

Xb

linear equivalent displacement of the bell [mm]

Xd

linear equivalent motion of the decoupler [mm]

k
X

state vector used in the Extended Kalman Filter

xxii

Nomenclature
XT

amplitude of mount base excitation [mm]

matrix of outputs for the least-squares method

vector of parameters in least-squares method

available displacement of decoupler plate [mm]

pressure differential [N/mm2 ]

Ps1s2 pressure differential between chambers in the test apparatus[N/mm2 ]


Ps1,s2

complex amplitude of the pressure differential in the test apparatus[N/mm2 ]

damping ratio

phase angle [ ]

density of fluid [kg/mm3 ]

18

constants used in nonlinear functions

xxiii

Chapter

Introduction
Automotive design encompasses many engineering disciplines and has captured the minds of
many individuals. Using advanced technologies, the vehicles of today have been designed to
represent a zone of comfort, handling, and driving pleasure.

Extensive effort is focused on

tuning the noise and vibration qualities to achieve the expected feel of a vehicle.
Reductions in body mass have increased engine-to-body weight ratios, which has raised
the level of noise, vibration and harshness within vehicles. More sophisticated hydraulic mounts
have been developed to address these issues. However, increased pressures to reduce vehicle
development cost has forced engineers to develop these complex hydraulic mount systems in a
limited time frame. The purpose of this work is to assist in reducing hydraulic mount development time by developing an effective hydraulic engine mount model and providing a physical
explanation of internal dynamics.

1.1

Desired Engine Mount Characteristics

Vehicle occupants receive undesirable vibrations through one of two possible excitation sources.
The first source, from engine eccentricity, typically contains frequencies in the range of 25 to
200 Hz with amplitudes generally less than 0.3 mm [20]. This frequency range corresponds to

1.1 Desired Engine Mount Characteristics

engine speeds of 750 to 6000 RPM for four cylinder engines. The second source of excitation
originates from road inputs and engine torque during harsh accelerations. Road inconsistencies
cause disturbances to the vehicle frame via the suspension system, whereas fierce accelerations
cause excessive engine torque and motion at the mounts. Excitations of this nature are typically
under 30 Hz and have amplitudes greater than 0.3 mm [20]. Each of these sources can be examined separately to demonstrate the desired mount characteristics, starting with the low frequency
excitations from road inputs.

1.1.1

Road Excitations

Low frequency, large amplitude excitations, transmitted from the road through the suspension
into the vehicle frame, can be represented as base excitation to the engine mount system.

To

demonstrate the desired mount characteristics, the system is simplified to a Single Degree-ofFreedom (SDOF) model illustrated in Figure 1.1a. Parameter m represents the engine mass,
while k and c denote the engine mount stiffness and damping characteristics, respectively. With
the disturbances from the vehicle body y(t), the equation of motion for this system is,
k(y x) + c(y x)
= m
x

(1.1)

This equation is converted to the frequency domain by assuming steady-state sinusoidal input and
output. Since the mount characteristics should minimize the relative displacement between the
engine and body, the frequency equations are manipulated to illustrate the relative displacement
transmissibility, given by (X Y )/Y . The equation becomes,

X Y
w
m
dr

=
2
Y
wdr m + wdr cj + k

(1.2)

where wdr is the driving frequency. Equation (1.2) is simplified using the nondimensional

damping coefficient = c/(2 mk) and plotted versus the nondimensional frequency ratio
p
r = wdr /wn , where wn = k/m in Figure 1.1b.

1.1 Desired Engine Mount Characteristics

x(t)

m
k

c
y(t)

Relative
Displacement Transmissibility Amp

X-Y
Y

z=0.1
5 desired
operating
range

z=0.3
z=0.6

0.5

1.5

2.5

3.5

4.5

Frequency Ratio (wdr/wn)

(a)

(b)

Figure 1.1: Low frequency excitations: (a) SDOF base excitation representation; (b) Frequency
domain relative displacement transmissibility.
Figure 1.1b represents the relative displacement transmissibility frequency domain plot.
This figure illustrates the desired operating frequency range for minimum engine to frame displacement. The driving frequency of base excitations should be below the systems natural frequency. Increasing the mount stiffness will yield higher natural frequencies and keep the large
amplitude vibrations within the displacement isolation region. Furthermore, increased damping will reduce the resonant amplitude and improve the relative displacements. Keeping these
characteristics in mind, the high frequency excitations will be examined next.

1.1.2

Engine Eccentricity

High frequency, low amplitude excitations, originating from engine eccentricity, can be represented as a force input to the mass of a fixed base SDOF system illustrated in Figure 1.2a. Forcing
on the engine mass Fo , represents the effective force input due to eccentricity and FT captures

1.2 The Hydraulic Engine Mount

transmitted force to the base. The resulting equation of motion for this system is
Fo = m
x + cx + kx

(1.3)

FT = cx + kx.

(1.4)

The ratio of transmitted force and input force,

FT
Fo

, represents the force transmissibility.

Using this relationship and the steady state solution in the frequency domain, the following
equation is developed.

FT
cwdr j + k

=
2
Fo
wdr m + cwdr j + k

(1.5)

As in the base excitation example, non-dimensional damping and frequency ratio are used
to plot the frequency response in Figure 1.2b. In this case, engine vibrations should operate
in the isolation region, above the natural frequency of the engine and mount system. Engine
mounts with low stiffness will reduce the rigid body natural frequency such that the operating
frequency of the engine eccentricities are within this region. Low damping characteristics will
further reduce the transmissibility.
The above analysis has determined the desirable characteristics of engine mounts. During low frequency, high amplitude vibrations, the ideal mount should exhibit large stiffness and
damping characteristics to reduce relative displacement transmissibility. Whereas for high frequency, low amplitude vibrations the ideal mount should have low stiffness and damping characteristics. These conflicting properties indicate that an ideal mount system has stiffness and
damping characteristics dependent on the amplitude of excitation. These amplitude dependent
characteristics are achieved in hydraulic engine mount designs.

1.2

The Hydraulic Engine Mount

The passive hydraulic mount, illustrated in Figure 1.3, consists of two fluid-filled chambers connected through a decoupler and inertia track. Typically, the fluid within the mount is a mixture

1.2 The Hydraulic Engine Mount

Fo(t)
m

x(t)

FT(t)

Force Transmissibility Amp (FT/FO)

z=0.1
desired
operating
range

z=0.3
z=0.6

0
0

0.5

1.5

2.5

3.5

Frequency Ratio (wdr/wn)

(a)

(b)

Figure 1.2: High frequency excitations: (a) SDOF forced input excitation; (b) transmitted force
frequency domain plot.
of ethylene glycol and water. The upper chamber is bound on top by the main rubber compliance
structure and on the bottom by a steel separator plate which houses the inertia track and decoupler. Under operating conditions the upper compliance holds the static weight of the engine and
as excitations are applied, a fluid pumping action occurs pushing fluid between chambers. The
separator plate, between the two chambers, is fixed to the base of the mount through a rigid
structure that surrounds the lower chamber. The lower chamber is bound by a compliant rubber
bellow, or lower compliance, that expands and contracts as fluid passes though the inertia track
and decoupler.
The decoupler plate has a finite travel distance within its cage. This device limits the
volume of fluid that can pass relatively freely between the upper and lower chambers. Once the
plate bottoms out on the cage the fluid must pass through the inertia track, a long column of fluid
between the two chambers. During small amplitude excitations the fluid passes freely through the

1.2 The Hydraulic Engine Mount

Upper
Compliance

Upper
Chamber

Decoupler
Decoupler
Cage

Lower
Chamber
Inertia
Track
Lower
Compliance

Figure 1.3: Cross section of a typical hydraulic engine mount.


decoupler into the more compliant lower chamber, giving the mount low damping and stiffness
characteristics. Fluid is forced through the inertia track during large amplitude vibrations, as
the decoupler bottoms out in its cage. The resistance and mass of fluid within the inertia track
generate increased stiffness and damping characteristics under these conditions.

1.2.1

Determining Mount Characteristics

Typically, hydraulic mounts are isolated and evaluated in the frequency domain using a servocontrolled hydraulic rate machine. Figure 1.5 illustrates the MTS 1000-Hz rate machine used
to determine the mount characteristics. During the test, the hydraulic actuator
1 is controlled
using a LVDT to excite the mount with a sine wave at a predetermined amplitude and frequency.
In addition, a static preload, or mean force, is applied to represent the static mass of the engine.
The transmitted force at the mount base is measured using a force transducer
2 . These readings
are used as feedback to control the mean force applied to the mount and to measure the amplitude
and frequency of the force applied during frequency excitation.

Using the MTS Teststar II

Dynamic Characterization software [22], the excitations are applied for a duration of 30 seconds

1.2 The Hydraulic Engine Mount

Servo-controlled
Actuator

Hydraulic
Mount
Force
Transducer

LVDT

Controller
Command
A/C Conditioner
D/C Conditioner
Data Acquisition

Workstation

Figure 1.4: Standard test configuration for hydraulic engine mounts.

Hydraulic
Actuator
1

Hydraulic
Mount
3

Force
Transducer
2

Figure 1.5: Illustration of MTS 1000-Hz rate machine used to measure dynamic characteristics.

1.2 The Hydraulic Engine Mount

for each frequency and amplitude condition. The program logic requires that each amplitude,
frequency and preload fall within a user described tolerance; 5% in all tests conducted during
this work. At the end of the frequency dwell, once the desired amplitude and frequency are
achieved, the high speed data acquisition samples both excitation X, and transmitted force FT ,
for a duration of time based on the frequency. Sine Regression is applied to each data sequence
to determine the amplitude and phase, as illustrated in Figure 1.6.
Imaginary

Aforce
Fforce
Real
Fexcitation
Aexcitation

Figure 1.6: Transmitted force and excitation vectors.

These vectors are used to calculate the cross point dynamic stiffness K and phase of the
mount, as

Af orce
Aexcitation

(1.6)

= f orce excitation

(1.7)

K =

The process is repeated for each test condition to develop the frequency domain system
characteristics. Between each frequency excitation the actuator is stopped for a mean dwell,
allowing the force to settle back to the constant preload Fp and resetting all internal dynamics of
the mount to zero.

1.2 The Hydraulic Engine Mount

1.2.2

Mount Characteristics

The decoupler gives the hydraulic mount its desired amplitude dependent characteristics; however, these mounts also exhibit liquid column resonance. Since the columns of fluid residing
between the inertia track and decoupler are not directly connected to the input excitation, they
generate additional degrees of freedom within the system. At some input frequency the liquid column, connected between the upper and lower chamber, will reach its natural frequency,
thus creating the liquid column resonance and altering the hydraulic mount characteristics. Figure 1.7a illustrates how the liquid column resonance of the inertia track influences the overall
dynamic characteristics of the fluid mount. This typically occurs between 5 to 25 Hz and causes

Phase (Deg)

Phase (Deg)

Dynamic Stiffness (N/mm)

Dynamic Stiffness (N/mm)

a rapid increase in the dynamic stiffness.

Frequency (Hz)

Frequency (Hz)

(a)

(b)

Figure 1.7: Typical mount dynamic characteristics: (a) Low frequency, large amplitude; (b)
High frequency, small amplitude.

Since there is also a mass of fluid associated with the decoupler, another liquid column
resonance in the high frequency range exists. Figure 1.7b, illustrates a typical high frequency
dynamic stiffness curve which includes the resonance of the decoupler. The high frequency

1.3 Literature Review

10

resonance of the decoupler mass is an undesirable effect, since it increases the dynamic stiffness
at high frequencies.
The characteristics of a hydraulic mount are more flexible and desirable than conventional
elastomeric mounts and their use in the automotive industry has become widespread. The trade
off for the enhanced flexibility is increased complexity within the hydraulic mount, and increased
design time. The following survey of papers establishes the current state of research in the area
of hydraulic mount modeling, which will clarify the focus of this thesis work.

1.3

Literature Review

The first paper revealing hydraulic engine mounts is by Bernuchon [1] in 1984.

The au-

thor presents the physical properties of a hydraulic mount and experimentally characterizes the
mounts dynamics. This paper is the first to introduce the excitation conditions and techniques
for dynamic stiffness and phase measurement. Simplified models are documented, but no mathematical development is given. Low frequency characteristics are related to liquid column resonance of the inertia track and high frequency experimental results are plotted up to 200 Hz. The
author then uses on-vehicle experimental results to conclude the superiority of hydraulic mounts
over conventional. In the same year, Corcoran and Ticks [4] describe hydraulic mounts and give
insight into the key characteristics. The influence of decoupler gap and inertia track geometry
is investigated using experimental dynamic stiffness and phase results from 1 to 40 Hz. A vehicle analysis using hydraulic engine mounts shows a 5dB improvement in the noise level at the
drivers ear for a 6 cylinder engine at 3300 RPM. The hydraulic mount is also shown to improve
engine bounce from road excitations. To conclude, the authors make recommendations on how
to apply hydraulic mounts to an existing vehicle.
Flower [5] builds on the understanding of hydraulic mounts by providing physical and
mechanical system models.

Bond graph techniques are then used to develop frequency do-

1.3 Literature Review

11

main mathematical expressions for the cross point dynamic stiffness. The models are the first
to associate an inertia with the decoupler, which is evident during high frequency low amplitude excitations. However, no experimental results are used to validate simulation results. In
the work by Ushijima et al. [32] the dynamic characteristics of a hydraulic mount are studied
in response to superimposed inputs. Experimental evidence is used to show that a compliant
decoupler provides improved performance to superimposed inputs by softening the strong decoupler nonlinearity. The authors developed a hydraulic mount which can achieve low dynamic
stiffness up to 800 Hz, by utilizing liquid column resonance and rubber surging. Also included
in the paper is an investigation into a semi-active mount incorporating electro-rheological fluids.
Seto et al. [29] provide a detailed analysis of the hydraulic mount and predict ideal performance
conditions. Design optimization theory for a dynamic absorber is applied to a mathematical
model of a hydraulic mount with only an inertia track. A simplified mechanical model is developed and the dynamic equations are derived analytically for use in optimization. A conditional
expression for minimizing engine vibration gives the optimum values of fluid mass and damping
of the inertia track. This work factors engine mass into the mount analysis and uses experimental results to validate the design method proposed. Further work focused on the inertia track is
conducted by Gau and Cotton [6]. The authors propose two categories of hydraulic mounts, single and double-load-bearing fluid chambers, and review mathematical models for both. A linear
lumped parameter model is developed for a standard single-load-bearing chamber mount with
only an inertia track, and results are compared to experimental data. The paper compares model
to experimental results in the low frequency range and investigates the influence of device parameters. The detailed study considers chamber compliance, inertia track area and length, and
fluid viscosity. A process for hydraulic mount design is presented, then a prototype fluid mount
is tuned to reduce vibration in an engine system.

1.3 Literature Review

12

The first of many papers by Singh et al. [30] represents a culmination of all work cited
to this point.

Linear, time-invariant, lumped parameter models for both free and fixed type

decouplers are mathematically developed. By splitting the hydraulic mount model into two, large
amplitude low frequency and small amplitude high frequency, the decoupler nonlinearity can be
avoided. The authors show that most prior models are special cases of their own. Mechanical
system models are developed, however they do not capture the transmitted force accurately.
Models are also presented in state-space form and validated over 1 to 50 Hz.

The authors

are the first to identify a problem with the hydraulic mount during high frequency excitations.
Experimental results show increased dynamic stiffness beyond 150 Hz and poor correlation with
the linear model, however the issue is not investigated further. The paper concludes with low
frequency parameter studies on the simplified model.
Branch and Haddow [20] provide a more fundamental explanation of the hydraulic mount,
using a previously developed linear model. This model is used to interpret the dynamic stiffness
curve by observing zeros and poles in the Laplace domain equation.

Low frequency model

validation is also conducted with the experimental data. A thorough review of engine vibration
sources and corresponding frequency ranges is also developed and used to explain desired engine
mount characteristics. Lee et al. [15] present modeling and performance analysis of a hydraulic
engine mount using the bond graph method. Model validation is shown over 1 to 50 Hz and 1 to
100 Hz ranges using experimental data.
More recent work begins to focus on the nonlinear dynamics of hydraulic mounts. Kim
and Singh begin an extensive nonlinear analysis of the hydraulic mount. In [12], the nonlinear
properties are identified for a fluid mount with an inertia track.

By experimentally measur-

ing the chamber volumetric compliance, an equation is fitted to the static curve, characterizing
a nonlinear chamber compliance. A nonlinear form of the fluid resistance within the inertia
track is also determined experimentally. A nonlinear lumped parameter model is formulated

1.3 Literature Review

13

and compared with experimental data over time and frequency domains from 1 to 50 Hz. Discrepancies between theory and experiment are attributed to gas-liquid phase transformation and
cavitation phenomenon. The content within their next two papers [11, 13] examines the dynamic
characteristics of a hydraulic mount in an isolated case and within a simplified vehicle model.
The nonlinear lumped parameter model is updated to include decoupler switching characteristics. Using the pressure differential and volume of fluid passing through the decoupler, flow is
switched by effectively increasing the pressure differential within the decoupler. Extensive experimental verification of the nonlinear model includes harmonic response data collected in the
time and frequency domain. To examine the vibration isolation and shock absorption properties,
simplified vehicle model experiments are conducted and include both harmonic and impulse excitations.

The mathematical model is validated for amplitude sensitive excitations from 1 to

50 Hz. However, the high frequency decoupler resonance shown experimentally is again not
pursued. The authors work also includes a new broadband adaptive hydraulic mount. Most
work documented in the previous three papers are a part of Kim [10].
Colgate et al. are the first to address the high frequency decoupler resonance [3]. Both
low and high frequency conditions are investigated as well as composite input response. The
authors conclude that high frequency isolation is significantly degraded when combined with a
large amplitude low frequency disturbance. Using two linear models for coupled and uncoupled
decoupler conditions, a piecewise linear simulation technique is used to show an amplitude dependent frequency response. State space and frequency domain equations for both linear models
are developed using the bond graph method. Model parameters are identified using the dynamic
stiffness curves of specially designed mounts, and the decoupler inertia is characterized using
high frequency experimental results. The piecewise decoupler model switches between integrating large and small amplitude equations depending on the displacement of the decoupler and
the decoupler velocity. Time domain simulations are conducted, however no results are shown.

1.3 Literature Review

14

The simulations represent experimental data over the complete low and high frequency ranges
of interest, however the authors state that moderate improvements can be attained by capturing
the effects of the lower chamber compliance, upper chamber bulge damping and leakage past
the decoupler. Margolis and Wilson [18] present a nonlinear model of the hydraulic mount that
uses first principle representation for inertia track resistance and decoupler. Momentum of the
decoupler is stopped by applying an additional resistive force, dependent on the position of the
decoupler and the pressure differential.

Simulation results for low frequency excitations are

validated using experimental data, however high frequency results do not correlate well.
The final set of papers places more emphasis on vehicle vibration with hydraulic mounts.
Ushijima and Dan [31] use the nonlinear building block approach to predict vibration of a vehicle
with hydraulic engine mounts. The authors combined the numerical simulation of a hydraulic
mount with modal analysis of a vehicle. In [8], Ishihama et al. present a new design method for
tuning low frequency vehicle vibration problems. The work uses engine and body modal analysis to determine mount contribution to vibration. Phase angle control on a hydraulic mount is
then used to adjust vibration levels within the vehicle. Inertia track damping is used to alter the
dynamic stiffness and phase angle at desired frequencies. The paper by Muller [23, 24] gives
design strategies for mount systems based on calculated and experimental methods. A rigid
body model of the engine and mount system is used to aid mount design for quasi-static and dynamic loads. FEA (Finite Element Analysis) tools are illustrated to be effective in predicting
axial stiffness, volumetric compliance and effective pumping area of a hydraulic mount. The authors present a standard and switchable hydraulic mount as well as a hydraulic bushing, but only
physical details and limited experimental results are shown. The strategies of transfer path analysis are explained and illustrated. Finally, the work investigates an active vibration absorber. In
the three papers by Royston and Singh [26, 27, 28], an extensive analytical analysis is developed
for nonlinear mounting systems based on the computational Galerkin method. Most of the work

1.4 Objectives and Thesis Overview

15

presented is focused on the response of a simplified engine and nonlinear mount system and all
nonlinear hydraulic mount modeling is based upon previous work by Kim and Singh. Little and
Kashani [16] also use existing models of hydraulic mounts to focus on decoupler switching characteristics. The authors apply smart materials such as magnetostrictive and electro-rheological
fluid to the flow communicating through the decoupler, to achieve semi-active and active hydraulic mounts.

1.4

Objectives and Thesis Overview

To reduce the time required to design a hydraulic mount, it is desirable to construct fluid mount
models that can be simulated to predict the system response before it is physically assembled.
From the preceding literature review, the shortcomings in hydraulic mount modeling are mainly
in the high frequency range. Colgate et al. [3] developed a model that is effective at high frequencies, however it is based on a piecewise linear approach.

No authors have presented a

physically intuitive model that is effective over the complete excitation range of interest. Furthermore, it is important for designers to understand how features within the hydraulic mount can
be changed to achieve the desired mount characteristics. Parameter investigations in the high
frequency range are also not presented in any literature reviewed.
This thesis consists of a complete guide to nonlinear hydraulic mount modeling and simulation, in an effort to reduce hydraulic mount development time.

Chapter 2 consists of an

extensive system modeling portion, that first redevelops the best models reviewed in the literature. Enhancements to the nonlinear decoupler model include a continuous function that follows
a simplistic, yet effective, approach to capture the switching effect. Leakage through the decoupler is considered and a clear nonlinear model of the transmitted force is shown. Upper chamber
bulge damping has also been accounted for in the system model. Resistance to oscillating fluid
flow is accounted for by using a simple equation that can be related to first principle fluid dy-

1.4 Objectives and Thesis Overview

16

namic equations. Finally, the modeling covers a unique bell component within the hydraulic
mount that alleviates the high frequency liquid column resonance of the decoupler.
Chapter 3 is dedicated to providing a physical explanation of low and high frequency characteristics. Time domain plots of the decoupler operation give an indication of how this component influences the system. Liquid column resonance is explained and related to system output.
In addition, the high frequency function of the additional bell component is presented in detail.
This portion also gives an indication of how each lumped parameter effects the overall system
characteristics.
Perhaps the most unique portion of this work is the experimental process, outlined in Chapter 4. This thesis covers a method whereby components of the hydraulic mount are isolated to
enable parameter identification. Chapter 4 introduces the method and apparatus, and documents
the calibration procedure.
Details of component parameter identification are given in Chapter 5. This includes the
test setup, perturbation conditions and system identification approach for each component.
The system simulation accounts for all experimentally determined nonlinearities within the
system, and is documented in Chapter 6. Experimentally measured parameters are incorporated
into a nonlinear simulation algorithm and compared with complete hydraulic mount system test
data.
Preliminary findings for parameter to geometry relationship are outlined in Chapter 7. This
brief section includes the initial findings for the relation between parameters identified and used
in the system simulation, and the geometry of the components.
Chapter 8 concludes the thesis. The key findings are highlighted and recommendations
for future work in this area are made.

Chapter

Mathematical Modeling of Hydraulic Engine Mounts


All mathematical models used in this thesis to represent hydraulic mount dynamic behavior are
developed in this chapter. First, a linear model of a typical hydraulic mount is derived and compared with previous work in this area.

Bond graphs are applied to extract system equations

while mechanical models are used to illustrate mount behavior. The chapter focuses on developing a nonlinear model that is effective across the frequency and amplitude ranges of interest;
the decoupler model represents an advancement in hydraulic mount modeling. Finally, a bell
system within the mount is presented and modeled. This additional component within a hydraulic mount is used to alleviate the high frequency resonance of the decoupler. To the best of
the authors knowledge, this system has not yet been modeled in any literature.

2.1

Linear Model of a Typical Hydraulic Mount

Linear models for a typical hydraulic mount are developed using the bond graph method. Lumped
parameters and variables are first assigned to physical characteristics, then the bond graph method
is applied to determine the linear system equations. The system equations are then divided into
two linear models, thereby providing a better representation of the actual system in both high
and low frequencies.

17

2.1 Linear Model of a Typical Hydraulic Mount

18

Upper
Compliance
C1 Ap kr br
Decoupler
Id R d

Upper
Chamber

Inertia
Track
Ii Ri

Lower
Chamber
Lower
Compliance
C2

Figure 2.1: Cross section of typical hydraulic mount.

2.1.1

Parameter and Variable Assignment

Consider the hydraulic mount cross section illustrated in Figure 2.1. The upper chamber compliance serves three main functions within the system.

First, it contains the main structural

stiffness and damping properties of the mount kr and br , respectively. The upper compliance
also functions as a piston, with an effective pumping area Ap . Finally, the rubber structure adds
volumetric compliance to the system, represented by C1 . The inertia track, a long column of
fluid between chambers, is assigned lumped parameters Ii and Ri representing the inertia and resistance, respectively. The other flow passage between the upper and lower chambers is through
the decoupler. To begin constructing a system model, the decoupler is assigned linear lumped
parameters Id and Rd , which also represent an effective inertia and resistance. Assigning the
decoupler an effective inertia is key to establishing high frequency characteristics and has been
included in [3, 5, 15, 32].

Finally, the lower chamber contributes to the volumetric compli-

ance and is modeled using a lumped parameter C2 . Using the assigned parameters, the system
schematic is converted to a lumped parameter model, illustrated in Figure 2.2.

2.1 Linear Model of a Typical Hydraulic Mount

19

X(t)
kr

br

P1(t)
Qi(t)

Qd(t)

C1
XT(t)

Ri Ii

Rd Id

FT(t)
P2(t)

C2

Ap

Figure 2.2: Lumped parameter system model.


Variables within the system include the input excitation X(t) and motion of the mount base
XT (t). Also, the upper and lower chamber pressures are captured by P1 (t) and P2 (t), respectively. Flow through the inertia track Qi (t) and decoupler Qd (t) are highlighted in Figure 2.2,
as well as the transmitted force to the mount base FT (t).

2.1.2

Bond Graph Model

The bond graph method is applied to develop equations of motion from the lumped parameter
model. This method maps the flow of energy within the hydraulic mount (see Appendix A or
van den Bosch et al. [2] and Karnopp et al. [25]) and is particularly useful for multi-disciplinary
systems. Since the hydraulic mount system is a combination of fluid and mechanical components, the bond graph method is applied as represented in Figure 2.3. This method has also been
applied in [3, 5, 15, 18] to determine the linear state equations.
The
Input excitation at the top of the mount is represented by a source of flow variable X.
effective piston area is now represented by the transformer variable and each compliance parameter is connected to pressure nodes of the upper and lower chamber. Variables and parameters in

2.1 Linear Model of a Typical Hydraulic Mount

20

:1/kr

: Ap

: X(t)

: br

: C1

: Ri

: Rd

: Qi(t)

: Qd(t)

: Id

: Ii

: C2

Figure 2.3: Bond graph model for a typical hydraulic engine mount.
the bond graph model represent those used in the lumped parameter system. However, since all
experimental studies are conducted with a fixed base test fixture, the base motion variable XT is
set to zero.

2.1.3

System Equations

Using the bond graph model, the equations for the internal dynamics of the system are established. The continuity equations become
C1 P1 = Ap X Qi Qd

(2.1)

C2 P2 = Qi + Qd

(2.2)

P1 P2 = Ii Q i + Ri Qi

(2.3)

P1 P2 = Id Q d + Rd Qd

(2.4)

and the momentum equations are

The system of four state variables and one input, represents the internal dynamics of a fluid
mount. The transmitted force developed using Figure 2.4 is dependent on the decoupler position,
however, a nonlinear function for this dependence will be derived later. Two situations for the

2.1 Linear Model of a Typical Hydraulic Mount


Ap

21
Ap

X(t)

X(t)
kr

br

P1(t)

Qd(t)

P2(t)

kr

br

Rd

FT(t)

P1(t)

Qd(t)
Rd

P2(t)
FT(t)
Ad

(a)

(b)

Figure 2.4: Transmitted force to the mount base: (a) decoupler closed; (b) free floating decoupler.
decoupler arise and can be considered separately, similar to the work in [3, 13]. For Figure 2.4a
the decoupler is contacting the cage and the transmitted force equation is

FT = kr X + br X + Ap (P1 P2 ) + Ap P2

(2.5)

and for Figure 2.4b where the decoupler is free

FT = kr X + br X + (Ap Ad )(P1 P2 ) + Ap P2 + Ad Rd Qd

(2.6)

In both cases the force transmitted to the mount base represents the sum of forces applied
through the mechanical stiffness and damping parameters, along with the influence of internal
chamber pressures.

In (2.5) the upper chamber pressure acts on the mount base and piston

area, while the lower chamber pressure is internal and does not influence FT . In (2.6) the upper
chamber pressure acts over the same area, however it is now reduced by the decoupler hole. Also,
the flow resistance through the decoupler influences the total transmitted force, captured initially
with Ad Rd Qd (similar to Colgate et al. [3]). Equations (2.1) to (2.6) can now be separated into
models that are dependent on the excitation amplitude; capturing the nonlinear behavior of the
hydraulic mount with two linear models.

2.1 Linear Model of a Typical Hydraulic Mount

2.1.4

22

Developing Two Linear Models

A linear approach to modeling hydraulic mounts can be used if small and large excitation amplitudes are considered as independent models, this follows a similar technique used by Singh et
al. [30] and Colgate et al. [3].
The Small Amplitude High Frequency Model assumes that during small amplitude excitations the decoupler does not contact the cage and all flow passes freely through the decoupler.
Also, the high frequency excitations are generally above the liquid column resonance of the inertia track, hence negligible flow passes through the inertia track. A linear model that captures the
system response for high frequency small amplitude excitations is developed by removing equation (2.3), and state variable Qi , from the state equations previously generated. The continuity
equations for the small amplitude high frequency become
C1 P1 = Ap X Qd

(2.7)

C2 P2 = Qd

(2.8)

and the momentum equation is


P1 P2 = Id Q d + Rd Qd

(2.9)

For this model, the decoupler is assumed to be free at all times, therefore Figure 2.4b applies and
equation (2.6) is used to represents the transmitted force.
FT = kr X + br X + (Ap Ad )(P1 P2 ) + Ap P2 + Ad Rd Qd .

(2.10)

The Large Amplitude Low Frequency Model assumes all flow passes through the inertia
track, since during large amplitude excitations the decoupler will spend the majority of time
against the cage with no flow passing through the decoupler. The equations for the linear model
are developed by removing the fluid momentum equation (2.4) from the system of equations

2.2 Mechanical System Model

23

(2.1)-(2.6) and by setting Qd to zero. The continuity equations become


C1 P1 = Ap X Qi

(2.11)

C2 P2 = Qi

(2.12)

and momentum equation is


P1 P2 = Ii Q i + Ri Qi

(2.13)

From Figure 2.4a, the decoupler is assumed fixed and thus (2.5) is the transmitted force equation.
FT = kr X + br X + Ap (P1 P2 ) + Ap P2

(2.14)

Using these linear system equations, equivalent mechanical system models can be developed to
illustrate the physical operation of the hydraulic mount.

2.2

Mechanical System Model

The linear system equations are now used to develop the mechanical model. The main area of
interest is the fluid column dynamics, since the stiffness and damping properties kr and br , are
already represented as a mechanical system. First consider the fluid system equations (2.7) to
(2.9) for the high frequency small amplitude model,
C1 P1 = Ap X Qd
C2 P2 = Qd
P1 P2 = Id Q d + Rd Qd
Substituting (2.7) and (2.8) into the derivative of (2.9) gives,
d + Rd Q d +
Id Q

1
1
+
C1 C2

Qd =

Ap
X
C1

(2.15)

2.2 Mechanical System Model

24

The system parameters and variables are converted to their equivalent linear mechanical form by
considering the ideal fluid system elements (see Rowell [25])

Qd = Ad X d
md
A2d
bd
=
A2d

(2.16)

Id =
Rd

In (2.16) md , bd , Ad and X d are the effective mass of the decoupler, the damping coefficient for
the linear decoupler motion, the area of the decoupler, and the linear velocity of the decoupler,
respectively. The compliance parameters are transformed from volumetric stiffness to a linear
stiffness using,
A2p
A2p
C1 =
, C2 =
k1
k2

(2.17)

where k1 and k2 are linear stiffness parameters of the upper and lower compliance, respectively.
Substituting (2.16) and (2.17) into (2.15) and integrating with respect to time, gives

d + bd X d + (k1 + k2 )
md X

A2d
Ad
X
=
k
X
d
1
A2p
Ap

(2.18)

with all initial conditions set to zero. Recall that (2.18) only represents the internal dynamics
of the fluid mount. To construct the complete mechanical system the additional stiffness and
damping variables of the upper chamber must be included. Figure 2.5 illustrates the mechanical
system model for the small amplitude high frequency model, following the same techniques as
Flower [5].
Before proceeding, the transmitted force to the base is derived for the mechanical model to
ensure this system is mathematically equivalent to the original system equations. Observing the
transmitted force at each pin in Figure 2.6, the force equations are determined.

2.2 Mechanical System Model

25

Figure 2.5: Mechanical representation of the high frequency, low amplitude linear model.

Figure 2.6: Transmitted force to the base in the mechanical model.

2.2 Mechanical System Model

26

FT _1 = kr X + br X
FT _2 = bd X d

(2.19)

Ad
Xd
Ap

Ad
Ad
d
= k1 X
Xd k2 Xd bd X d md X
Ap
Ap

FT _3 = k2
FT _4

These equations are then rewritten in terms of the fluid system parameters using the transformations in (2.16) and (2.17) and also using the relationships in (2.7) to (2.9):
FT _1 = kr X + br X
FT _2 = Rd Ad Qd

(2.20)

FT _3 = Ap P2
FT _4 = (Ap Ad ) (P1 P2 )
The summation of the transmitted force equations in (2.20) is
FT = kr X + br X + (Ap Ad ) (P1 P2 ) + Ap P2 + Ad Rd Qd

(2.21)

Comparing equation (2.6) with (2.21) validates that Figure 2.5 corresponds to the small amplitude high frequency model. Similar work was presented by Flower [5], however no mathematical
validation was given.
The next logical extension is to incorporate the inertia track into the mechanical model.
However, with both the decoupler and inertia track the model becomes nonlinear. To illustrate
the concept, Figure 2.7 includes an inertia track mass with a switch that engages the inertia track
when the decoupler motion reaches a certain displacement. This approach has also been presented in [5]. Also, the damping element on the inertia track is not connected to the ground
where transmitted force is calculated since the inertia track travels perpendicular to the direction
of transmitted force. The nonlinear mechanical model is developed solely for illustrative pur-

2.3 Nonlinear Model Extension

27

Figure 2.7: Mechanical model that includes the decoupler and inertia track.
poses and is not studied any further in this thesis. Instead, effort will be placed on enhancing the
fluid system model to capture nonlinear effects.

2.3

Nonlinear Model Extension

This portion of the hydraulic mount modeling is devoted to the development of a nonlinear
lumped parameter model that will represent the mount response over the complete excitation
range. The nonlinear model evolves by enhancing the parameters of the linear model developed
in Section 2.1. These enhancements capture nonlinearities in the upper chamber stiffness, effective pumping area, chamber volumetric compliance, inertia track resistance, decoupler switching,
transmitted force, and decoupler leak flow.
The model enhancements have been developed in conjunction with component evaluations,
conducted experimentally and documented in Chapter 5. To the best of the authors knowledge,
this section represents an advancement to the nonlinear modeling work of Kim and Singh [11, 13]
and Colgate et al. [3].

2.3 Nonlinear Model Extension

2.3.1

28

Upper Chamber

Stiffness and Damping In the preceding sections the variables kr and br have been introduced to
represent the stiffness and damping properties within the upper chamber. In all papers reviewed,
these variables are treated using the Viogt model for rubber, which assumes both properties are
invariant with frequency and amplitude. In an effort to develop a complete non-linear model,
more attention has been given to associating the variability of these parameters with the frequency and amplitude of excitation. Since methods are unavailable to reliably predict dynamic
characteristics of rubber from physical geometry, the characteristics will be gathered experimentally and identified in Chapter 5. The behavior of system parameters kr and br in the frequency
domain are shown to be dependent on the steady-state amplitude and frequency of the excitation,
as well as the preload force magnitude. Stiffness and damping are written in the functional form

Fp
kr wdr , X,

(2.22)

Fp )
br (wdr , X,

the amplitude, and Fp is the preload force.


where wdr represents a frequency of oscillation, X
Effective Pumping Area The effective pumping area is dependent on the geometry of
the rubber, but the parameter cannot be easily determined from the geometry. Under the static
load of the engine mass, the rubber is deformed prior to the excitations. This preload force
Fp yields a mean displacement of the rubber that influences the effective pumping area. The
effective pumping area Ap is a function of the preload displacement and has been experimentally
measured in Chapter 5. Since the preload displacement is directly related to the static force,
through the static load displacement curve, the effective pumping area is written as a function of
the preload force

Ap (Fp )

(2.23)

2.3 Nonlinear Model Extension

29

Compliance The variable C1 represents the volumetric compliance of the upper chamber,
however it does not capture the hysteresis that is evident in the volumetric expansion. Hysteresis
is a result of molecular chains unraveling as the material expands and contracts, and is shown
experimentally to be evident in the expansion of the upper chamber in Chapter 5. The net effect
on the system response is an apparent increase in damping.

Colgate et al. [3] also suggest

that models are moderately improved by including bulge damping effects that are the result of
hysteresis. In the fluid system representation the volumetric damping is captured by including a
resistance parameter on the flow into the compliant region.
Figure 2.8 illustrates the changes from the standard modeling technique.

Figure 2.8a

shows a fluid system representation of the upper chamber compliance used in previous models,
while Figure 2.8b illustrates how the bulge damping has been modeled with the addition of a
flow resistance. A bond graph of this model is illustrated in Figure 2.9 and is used to develop
the equation
C1 P1 = QT + C1 R1 Q T

P1(t)

(2.24)

P1(t)
C1

Qt(t)

C1

Qt(t)
R1

(a)

(b)

Figure 2.8: Capturing bulge damping in the upper chamber: (a) previous model; (b) addition
of a resistance term in the compliance.

Here, QT has been introduced to represent flow entering into the compliant region and the
parameter R1 represents the flow resistance. The variable P1 is the same upper chamber pressure
used in previous models.

2.3 Nonlinear Model Extension

30
:R1

QT(t)

:P1(t)

: C1

Figure 2.9: Bond graph model of the compliance system with resistive parameter.
The same issues arise with the parameters C1 and R1 , as with the mechanical stiffness and
damping properties. These parameters are dependent on the steady-state volume amplitude VT
and the frequency wdr of excitations. Due to the geometrical changes under static preload, the
nonlinear parameters are also considered as functions of the preload force Fp
C1 (wdr , VT , Fp )

(2.25)

R1 (wdr , VT , Fp )
These parameters are determined experimentally in Chapter 5.

2.3.2

Lower Chamber

The influence of the lower chamber compliance on the mount characteristics is, to a large extent,
dependent on the system design. Hydraulic mounts that use the lower chamber as a load bearing
chamber will generally have stiff lower compliance values that significantly influence the system dynamics. If this is the case, the volumetric parameters should be assigned to capture the
compliance and bulge damping characteristics, similar to (2.25). However, if the main function
of the lower chamber is to accommodate fluid transfer, the volumetric compliance will have little influence on the system dynamics. Under these conditions a single volumetric compliance
parameter C2 is used to characterize the lower compliance. This design approach is the most
common and is the one considered in this work.

2.3 Nonlinear Model Extension

31

Under certain circumstances the volume of fluid pushed into the lower chamber during
preload will increase the chamber compliance parameter as the rubber stretches. Depending
on the design, it may be appropriate to model the lower chamber compliance as a function of
the static preload displacement, which is related to the preload force through the static load
displacement curve,

C2 (Fp )

2.3.3

(2.26)

Inertia Track

The linear inertia track model assumes a constant inertia Ii and resistance Ri . To enhance this
model, some fluid dynamic properties are considered. Kim [10] has concluded that the fluid
density can be considered constant and that the thermodynamic effects on the flow are negligible
for the oscillatory flow in an inertia track. However, Kim and Singh [13] have used an increased
fluid inertia for laminar flow within the inertia track. All other papers regarding the inertia track
dynamics assume a constant inertia parameter. Using previous research and experimental results
detailed later, the inertia parameter is concluded to be a constant parameter in the model.
Unlike the inertia parameter, the resistance does not exhibit constant behavior. To determine the flow resistance trends within the inertia track, this work first considers the fluid
equations for losses through a pipe (see White [33, pg.363]) by

P = f

8L 2
8 2
Q
+
k
Q
2 d5
d4

(2.27)

where L and d are the length, and diameter of the pipe and is the fluid density. Also,

f d , RED and k (d, D) represent friction factor coefficients for a pipe and entrance losses, respectively. These coefficients are dependent on the Reynolds number (RED ), the mean surface

roughness height , and reservoir diameter D.

2.3 Nonlinear Model Extension

32

Equation (2.27) can be used to model continuous fluid flow into a pipe with some reasonable degree of accuracy, however the inertia track flow is oscillatory. Since the flow is sinusoidal,
the Reynolds number will change at each point within a period of flow oscillation. As the flow
is changing direction the Reynolds number will approach zero, then it will increase to a maximum, passing through the laminar and into the turbulent region. The amount of time spent in
the laminar and turbulent regions will be dependent on the amplitude and frequency of the flow.
Since both f and k will fluctuate as the flow oscillates, a mean value for each is used. The flow
variable Q, now written as the inertia track flow variable Qi , is factored out of (2.27) and an effective resistance value Ri2 is defined. This is combined with the existing resistance parameter
Ri in (2.3), now called Ri1 , to form the nonlinear resistance equation

P = (Ri1 + Ri2 |Qi |) Qi

(2.28)

The inertia term is now combined to provide a nonlinear fluid momentum equation from (2.3).
P1 P2 = Ii Q i + (Ri1 + Ri2 |Qi |) Qi

(2.29)

In (2.28) and (2.29) it is necessary to place the flow variable Qi inside absolute value signs
to achieve the appropriate influence on the pressure differential.

2.3.4

Decoupler

Flow Control The nonlinear switching behavior of the decoupler accounts for most of the amplitude dependent characteristics.

When flow is oscillating across the decoupler orifice and

the decoupler plate does not contact the cage, the decoupler is considered open and the system
behaves as a simple orifice. In equation (2.4) the decoupler is assigned a lumped inertia and resistance value similar to the inertia track. The inertia of the decoupler exists, and is considered
to include the decoupler plate and a column of fluid above and below. As with the inertia track,

2.3 Nonlinear Model Extension

33

the decoupler inertia is assumed constant with amplitude and frequency of flow excitation and is
experimentally validated later.
The free-floating decoupler resistance is analyzed in a similar fashion to the inertia track.
Starting with the equations for fluid loss through an orifice (see White [33, pg.363])
P =

2
Q
A2d

(2.30)

where is the fluid density and Ad is the decoupler orifice area. Also, (, RED ) is an orifice
friction factor which is a function of the Reynolds number RED and the velocity-of-approach
factor . Again, the flow is oscillatory and thus the Reynolds number does not remain constant.
Using the same oscillatory flow rationale used for the inertia track, the nonlinear momentum
equation for the free floating decoupler becomes
P1 P2 = Id Q d + (Rd1 + Rd2 |Qd |) Qd

(2.31)

where the parameters Rd1 and Rd2 represent the new resistance parameters.
Equation (2.31) applies to the case where the decoupler plate is free from contacting its
cage. When the decoupler plate does contact the cage it blocks all flow across the decoupler
and introduces the amplitude dependent nonlinear characteristics. The additional flow-stopping
effect can be mathematically represented by making the decoupler orifice area approach zero,
Ad 0, in equation (2.30), which drives Qd 0.

This effect is equivalent to adding a

relatively large resistance term to equation (2.31). The additional resistance term is captured
with Radd and is applied to equation (2.31) to give
P1 P2 = Id Q d + (Rd1 + Rd2 |Qd | + Radd ) Qd

(2.32)

A similar approach has been taken in [11, 13, 18].


Since the flow stopping resistance Radd is a nonlinear parameter, the switching characteristics must be established. In [11, 13, 18] the decoupler switching characteristics are dependent
on the volume of fluid passed through the decoupler and the pressure differential between upper

2.3 Nonlinear Model Extension

34

and lower chambers. However, the work for this thesis has established that the switching characteristics are dependent on the volume of fluid and the direction of flow across the decoupler.
During low frequency excitations this differentiation will have no influence on the nonlinear response, since the low inertia decoupler column moves in phase with the pressure differential.
However, at high frequencies, typically above 100 Hz, the inertia effect will cause the decoupler
column to oscillate out of phase with the pressure differential.
Using the volume of fluid and the flow direction across the decoupler as switching variables, a mathematical representation is now established. A nonlinear function for Radd is developed using the decoupler position in an exponential function with the arctangent function to
model the flow dependent behavior.
The flow of fluid through the decoupler Qd is integrated to obtain the decoupler volume
Z T
Vd =
Qd dt
0

with the decoupler initially positioned on centre at time zero. The decoupler position Xd is
related to the volume via the measured decoupler area Ad .
Xd =

Vd
Ad

(2.33)

As the decoupler contacts the cage limits, the resistance Radd takes effect and reduces
the decoupler flow to effectively zero. When the flow reverses direction, a result of the input
pressure differential and momentum within the system, the additional resistance is removed as
the decoupler travels relatively freely to the opposing cage limits. Figure 2.10 illustrates the
decoupler sign convention and the desired function of Radd .
An exponential function is developed to model the desired additional resistance term or
flow-stopping action shown in Figure 2.10.

Using the decoupler position (2.33) and a new

switching parameter Rd3 , the additional resistance term becomes


Radd = 1 eRd3 Xd arctan(Qd 2 )

(2.34)

2.3 Nonlinear Model Extension

35

Radd
-Qd(t)

+Qd(t)

+Qd(t)

-Qd(t)

Qd(t)
0

Xd

Xd
(a)

(b)

Figure 2.10: The nonlinear decoupler resistance: (a) sign convention; (b) switching function.
where 1 is a constant used to place the overall additional resistance magnitude in the appropriate
range.

To achieve negligible resistance when the decoupler is free, is assigned a constant

value of 1 1016 for all cases.

The parameter Rd3 is introduced to control the decoupler

position at which the exponential increases and the additional resistance reduces decoupler flow.
Figure 2.11 indicates how an increase in Rd3 is comparable to closing the decoupler gap.
Flow dependent characteristics of equation (2.34) are developed by switching the exponential sign using the principle branch of the arctangent function (arctan). This function is applied
to the decoupler flow to achieve a continuous output of 1, dependent on the direction. The
constant 2 is included to adjust the shape of the function output. To demonstrate how 2 was
selected, consider the arctangent function in isolation

2
arctan(Qd 2 )

(2.35)

Figure 2.12 illustrates the impact of constant 2 on the shape of this function. To achieve a crisp
switching response the constant 2 is set to 1 105 s/mm3 for all cases. The

in (2.35) is

included in the parameter Rd3 of equation (2.34) to simplify the expression.


The final form of the momentum equation (2.31) becomes

P1 P2 = Id Q d + Rd1 + Rd2 |Qd | + 1 eRd3 Xd arctan(Qd 2 ) Qd

(2.36)

2.3 Nonlinear Model Extension

36

-7

Additional Resistance (kg/s-mm )

x 10

4.5
4
3.5
3
2.5
2
1.5
1
0.5

-0.45

-0.30

-0.15

0.15

0.30

0.45

Decoupler Position (mm)


Rd3 = 35

Rd3 = 32

Rd3 = 28

Figure 2.11: Additional resistance function indicating influence of parameter.

1
0.8
0.6
0.4
0.2
0
-0.2
-0.4
-0.6
-0.8
-1
-100

-80

-60

-40

-20

20

40

60

80

100

x2 = 1x105

Decoupler Flow (mm )


x2 = 1

x2 = 1x10-1

Figure 2.12: Illustration of the arctangent function.

2.3 Nonlinear Model Extension

37

This equation represents a nonlinear function that captures the decoupler momentum and introduces one new parameter to the model Rd3 . Equation (2.36) is also used to identify Rd3 from
experimental data, described in Chapter 5.
The transmitted force equations are also influenced by the decoupler system and will be
considered next.
Transmitted Force Improvements in the transmitted force equations (2.5) and (2.6) are
established by including the effects of the cage geometry and switching properties. The transmitted force equations for a typical decoupler and cage model are developed first. Then the
transmitted force equations for the open and closed decoupler situations are combined into one
function through the development of a nonlinear switching function.
The influence of the decoupler cage on the transmitted force equations is modeled using
the uncoupled isolated decoupler system shown in Figure 2.13. The cage geometry includes
the area of the holes Ad_holes and the area of the decoupler plate Ad , while the total area of the
holding plate is denoted by Ap . The pressure differential across the system is represented by the
nonlinear equation (2.36). For the uncoupled case the additional resistance function becomes
negligible and the momentum equation becomes

P1 P2 = Id Q d + (Rd1 + Rd2 |Qd |) Qd

(2.37)

As Figure 2.13b illustrates, the pressures P1 and P2 are introduced as intermediate pressures within the decoupler and are used to establish the regions of pressure change

P1 = P1 P1
0

Pplate = P1 P2
0

P2 = P2 P2

(2.38)

2.3 Nonlinear Model Extension


Ad_holes

P1(t)

38
Qd(t)

P 1(t)

P 2(t)
Ad

P2(t)

Qd(t)

FD

Ap
(a)

(b)

Figure 2.13: Isolated decoupler model that includes cage geometry: (a) geometry parameters;
and (b) pressure gradient breakdown.
where P1 and P2 represent pressure drops across the cage and Pplate denotes the pressure
differential across the decoupler plate. The total pressure across the system is
P1 P2 = P1 + Pplate + P2

(2.39)

Following the lumped parameter approach, the pressure changes P1 and P2 are considered to be only a result of flow through the cage holes and some lumped resistance Rcage . Also,
Pplate is thought to represents a pressure change due to the inertia within the decoupler Id . This
approach assigns the momentum equation to distinct portions of the pressure differential, as
P1 = Rcage Qd
Pplate = Id Q d
P2 = Rcage Qd

(2.40)
(2.41)
(2.42)

Since the resistance across the top and bottom cage are the only contributors to the total
decoupler resistance, P1 and P2 are summed and related to the resistance in equation (2.37).
P1 + P2 = (Rd1 + Rd2 |Qd |) Qd

(2.43)

Substituting (2.41) and (2.43) into equation (2.39) yields the same pressure relationship
established in (2.37), therefore this procedure has not changed the system dynamics.

2.3 Nonlinear Model Extension

39

By splitting the decoupler pressure differential, the transmitted force is established using a
force balance. From Figure 2.13b
FD = (Ap Ad )(P1 P2 ) + (Ad Ad_holes )P1 + (Ad Ad_holes )P2

(2.44)

and applying the relationship in (2.43) to equation (2.44) gives,


FD = (Ap Ad )(P1 P2 ) + (Ad Ad_holes ) (Rd1 + Rd2 |Qd |) Qd

(2.45)

If the upper chamber stiffness and damping are included, along with the lower chamber pressure,
the total transmitted force becomes
FT = kr X + br X + (Ap Ad )(P1 P2 ) + Ap P2

(2.46)

+(Ad Ad_holes ) (Rd1 + Rd2 |Qd |) Qd


This equation is developed using the model that assumes the decoupler is open, or free
from contacting the cage. When the decoupler contacts the cage, the transmitted force becomes
equivalent to
FT = kr X + br X + Ap (P1 P2 ) + Ap P2

(2.47)

matching equation (2.5) in the linear model.


Two events occur to transform the transmitted force equations from (2.46) to (2.47). First,
the decoupler closes and the pressure differential acts over the full area Ap , essentially Ad 0.
Second, flow across the decoupler becomes negligible, Qd 0.

Since the flow across the

decoupler is already controlled via the nonlinear decoupler resistance previously developed, only
the decoupler area function still needs to be developed.
The decoupler area is dependent on the decoupler position and pressure differential, following [11, 13, 18]. When the decoupler plate is contacting the lower cage and the pressure
differential P = P1 P2 is positive, Ad will appear to be zero. However, if the pressure differential reverses direction the decoupler orifice will appear open, even if the inertia still holds the
decoupler in place. These switching conditions are different from those used in the flow control

2.3 Nonlinear Model Extension

40

Ad_fnc
+DP(t)

-DP(t)

Ad

P1(t)
-Vd_max
0

P2(t)

Vd_max

-DP(t)
-Vd_max

(a)

+DP(t)
0

Vd_max

Vd

(b)

Figure 2.14: The decoupler area: (a) sign convention; (b) switching function.
since the transmitted force is directly related to the pressure differential across, not flow through,
the decoupler.
In the following nonlinear development the parameter Vd_ max is introduced to signify the
maximum volume of fluid accommodated by the decoupler. Using the sign convention illustrated
in Figure 2.14a, the nonlinear Ad_f nc function is illustrated in Figure 2.14b. The conditions for
the decoupler area switching are

Ad_f nc =

Ad , Vd_ max < Vd < Vd_ max


0,
Vd Vd_ max ,
Ad ,
Vd Vd_ max ,
0,
Vd Vd_ max ,
Ad ,
Vd Vd_ max ,

P
P
P
P

>0
0
<0
0

(2.48)

To develop a continuous function for Ad_f nc the arctangent function is used on both the

pressure differential and decoupler volume displacement. The function development first includes an arctangent function of the pressure differential multiplied by the volume displacement.
A=

2
Vd arctan ( 3 P )

(2.49)

The constant 3 controls the shape of the arctangent function similar to equation (2.35) and is set
to 1 105 mm2 /N. For positive pressure differentials, the value of A will become Vd . So, when
the pressure differential is negative and the volume displacement is up near the top cage, Vd is
also negative and the value of A will again be positive. By subtracting Vd_ max from A and taking

2.3 Nonlinear Model Extension

41

the arctangent of the result, the function will only give a result of +1 when A exceeds Vd_ max .
The function becomes
B=

2
arctan ((A Vd_ max ) 4 )

(2.50)

where a constant 4 is again used to control the shape of the arctangent function, with a value
of 1 103 mm3 . From the conditions on Ad_f nc outlined in (2.48), the value of B will be
+1 for each case where Ad_f nc = 0 is desired and B will take on 1 for each case where
Ad_f nc = Ad is desired. The final step of this function development is to establish an equation
for this relationship.
1
Ad_f nc = Ad (1 B)
2
Writing the nonlinear equation in terms of the pressure differential and volume, gives

2
Ad_f nc = Ad
arctan
Vd arctan ( 3 P ) Vd_ max 4

(2.51)

(2.52)

Finally, the complete nonlinear transmitted force equation from equation (2.46) is
FT = kr X + br X + (Ap Ad_f nc )(P1 P2 ) + Ap P2 +

(2.53)

(Ad Ad_holes ) (Rd1 + Rd2 |Qd |) Qd


At this point the nonlinear analysis has captured the dominant switching characteristics
of the decoupler, but an adjustment must be made in the cases where the decoupler does not
completely block the flow.
Leak Flow The nonlinear, decoupler flow control equation completely stops decoupler
flow.

However, some decoupler designs permit high resistance flow after the decoupler has

contacted the cage. The leak flow passes between the decoupler and cage when the decoupler
is in the closed position, thus it only exists when Qd is negligible. For convenience this flow is
referred to as decoupler leak flow, represented by the variable QL . The current decoupler model
is enhanced to capture this leak flow with the addition of an orifice, as illustrated in Figure 2.15a.

2.3 Nonlinear Model Extension

42
RL2_fnc

+Qd(t)

-Vd_max

-Qd(t)

+Qd(t)
RL2

QL(t)

Qd(t)

Vd_max

-Qd(t)

RL2 x x6

-Vd_max

(a)

Vd_max

Vd

(b)

Figure 2.15: Decoupler leak flow: (a) system model and sign convention; (b) nonlinear resistance function.
Using equation (2.30) for losses through an orifice, the resistance on the leak flow is written as
P1 P2 = RL2 |QL | QL

(2.54)

where RL2 represents a simplified resistance term.


It should be noted that the leak flow has not been modeled as high resistance flow through
the decoupler. Since the decoupler flow is integrated to attain the position, it is important to
preserve the memory of the decoupler plate by completely stopping the decoupler flow. There
are, however, two approaches to modeling the resistance to leak flow. The first is to consider
the lumped parameter RL2 as a constant and assume that during free decoupler motions negligible flow will pass through the leak orifice. The second approach is to develop a variable
resistance orifice which only exists when the decoupler is contacting the cage and the flow is in
the appropriate direction. In the following, the variable resistance function is developed.
Conditions for the nonlinear resistance function are

, Vd_ max < Vd < Vd_ max

Vd Vd_ max ,
RL2,
,
Vd Vd_ max ,
RL2_f nc =

RL2,
Vd Vd_ max ,

,
Vd Vd_ max ,

Qd
Qd
Qd
Qd

>0
0
<0
0

(2.55)

Based on these conditions a continuous function is developed, as illustrated in Figure 2.15b.


The arctangent function is used again to provide the necessary continuous switching properties,

2.3 Nonlinear Model Extension

43

similar to the previous development of Ad_f nc . Starting with equation (2.50), the pressure differential is now replaced with the flow through the decoupler to give
2
C = arctan


2
Vd arctan( 5 Qd ) Vd_ max 6

(2.56)

where 5 and 6 are used to control the shape of the arctangent function, both with values of
1 105 . The value of C will be +1 for each case where Qd is in the positive direction and Vd
has exceeded Vd_ max or when Qd is in the negative direction and Vd is less than Vd_ max . All
other combinations of Qd and Vd will yield a C value of 1. From the nonlinear conditions of
RL2_f nc in equation (2.55), when C is +1 the desired value of RL2_f nc is RL2 , and when C is 1
the value of RL2_f nc should be . A function that provides this relationship is
RL2_f nc = RL2 (1 + ( 7 7 C))

(2.57)

where the value of 7 is the multiplicative factor that numerically approaches , and is 1 103
in this model. The complete nonlinear function for RL2_f nc using (2.56) and (2.57) is

RL2_f nc

2
2
= RL2 1 + 7 7 arctan
Vd arctan( 5 Qd ) Vd_ max 6

(2.58)

and equation (2.54) becomes

P1 P2 = RL2_f nc |QL | QL

(2.59)

Although not every decoupler design will exhibit leak flow, this particular development
will be used to model the mount studied in Chapters 5 and 6.
The final portion of this chapter is focused on developing a model that represents the hydraulic mount with a bell system.

2.4 Modeling the Additional Bell System

44

Bell
Chamber
Bell
Upper
Chamber

Inertia
Track
Lower
Chamber

Decoupler

Figure 2.16: Cross section of a hydraulic mount with a bell system.

2.4

Modeling the Additional Bell System

As illustrated in Figure 1.7b, the high frequency resonance of the decoupler creates an undesirable increase in the dynamic stiffness characteristics. A new design feature has been developed
that adds more flexibility to the high frequency results and alleviates the decoupler resonance
problem. The concept is to place an additional plate, or bell, within the upper chamber and
attached directly to the excitation point. The bell splits the upper chamber in two and creates
another flow passage, as illustrated in Figure 2.16. With the additional flow passage the system
takes on an another degree of freedom. In the following, a lumped parameter fluid system model
will be developed and the nonlinearity issues discussed. Also, a mechanical system representation is presented.

2.4.1

Lumped Parameter Assignment

A lumped parameter model of the hydraulic mount shown in Figure 2.16 is illustrated in Figure 2.17. As seen in both Figures 2.16 and 2.17, the upper chamber is now divided in two with
the volume above the bell plate now referred to as the bell chamber and assigned the compliance

2.4 Modeling the Additional Bell System

45

X(t)
Ap
Rb Ib
kr

br

Pb(t)

Ab
Qb(t)
Qi(t)

Cb

P1(t)
Qd(t)

C1
XT(t)

Ri Ii

Rd Id

FT(t)
P2(t)

C2

Am

Figure 2.17: Lumped parameter fluid system model including the bell.
parameter Cb . The chamber below the bell and above the decoupler and inertia track assembly
is still referred to as the upper chamber, keeping the same compliance parameter C1 . Above the
bell the effective pumping area is still denoted using parameter Ap . However, the area of the
mount at the bell is now captured with Am . The gap between the bell and mount wall can be
considered as an annular orifice with an area denoted by Ab . Under oscillating conditions, fluid
flow through area Ab is assigned lumped parameters Ib and Rb to capture the inertia and resistance, respectively. The inertia parameter represents a column of fluid communicating between
the bell chamber and the upper chamber. Although this inertia is quite low, it introduces an additional degree of freedom within the system which becomes apparent in the high frequency range.
The inertia track, decoupler and lower chamber receive the same lumped parameters used in the
original model.
Additional variables within the system include the pressure in the bell chamber Pb (t), and
the flow into the bell chamber Qb (t).

2.4 Modeling the Additional Bell System


: -(Am-Ab-Ap)

: br

:Pb(t)

46
: Cb

: Rb

.
X(t)

: Qb(t)

: Ab

: 1/kr

Am-Ab

:P1(t)

: Ib

: C1

: Ii

: Id

: Qi(t)

: Qd(t)

: Ri

: Rd
: P2(t)

: C2

Figure 2.18: Bond graph model that includes the inertia track, decoupler and bell.

2.4.2

Bond Graph Model

Using the lumped parameter fluid system model in Figure 2.17, the bond graph method is applied
to extract the system equations. Figure 2.18 represents a bond graph model for a hydraulic mount
with the bell system. Displacement excitation at the top of the mount is represented by a source
Three transformers are used to communicate between the linear motion and
of flow variable X.
displacements of fluid within the system. Each volumetric compliance parameter is connected
to the pressure nodes of the bell, upper and lower chamber; while, each flow variable resides
between the corresponding chamber pressures. Parameters and variables associated with the
bond graph model follow the same notation as in Figure 2.17. Since all experimental studies are
conducted with a fixed base test fixture, the motion of the base XT is set to zero.

2.4 Modeling the Additional Bell System

2.4.3

47

System Equations

From the bond graph model, the continuity equations for each of the chambers are
Cb P b = Qb (Am Ab Ap )X

(2.60)

C1 P1 = (Am Ab )X Qb Qd Qi

(2.61)

C2 P2 = Qd + Qi

(2.62)

The momentum equation for each fluid system is

P1 Pb = Ib Q b + Rb (Qb + Ab X)

(2.63)

P1 P2 = Id Q d + Rd Qd

(2.64)

P1 P2 = Ii Q i + Ri Qi

(2.65)

These equations capture the internal dynamics of the hydraulic mount with a bell in the upper chamber, however they do not include the output force equation. The transmitted force to
the mount base is determined using the same technique as used for equation (2.6). As previously
determined, the transmitted force equation developed from the high frequency linear model captures all the desired characteristics and can be extended to the low frequency and nonlinear model
easily. Using Figure 2.17 and the linear form in (2.6), the transmitted linear force equation for
the free decoupler case is
FT = kr X + br X + (Am Ad )(P1 P2 ) + Am P2 + Ad Rd Qd (Am Ap )Pb

(2.66)

The additional term in (2.66) (Am Ap )Pb , represents the force exerted on the base from the bell
chamber pressure by ignoring the bell gap Ab . Equation (2.66) can be enhanced to a nonlinear
form with (2.53).
The set of equations (2.60) to (2.65) are of the same form as the linear equations first
presented for the standard hydraulic mount. This model can be handled similarly to the previous

2.5 Mechanical System Model with a Bell

48

cases, either by breaking the equations into two linear equations for low and high frequency
excitations, or by incorporating the nonlinear decoupler model.

2.5

Mechanical System Model with a Bell

The mechanical system model that includes the bell is developed using techniques similar to
thoes presented in the developments of the mechanical model for a standard hydraulic mount.
Since the mechanical system must be developed using linear system equations, the equations (2.60)
to (2.65) are considered for high frequency excitations only. The high frequency linear model
including just the decoupler and bell dynamics and becomes
Cb P b = Qb (Am Ab Ap )X

(2.67)

C1 P1 = (Am Ab )X Qb Qd

(2.68)

C2 P2 = Qd

(2.69)

P1 Pb = Ib Q b + Rb (Qb + Ab X)

(2.70)

P1 P2 = Id Q d + Rd Qd

(2.71)

Substituting equations (2.67) to (2.69) into the derivative of (2.70) and (2.71), gives the two
coupled liquid column system equations

1
1
1

Ib Qb + Rb Qb +
+
Qb +
Qd =
C1 Cb
C1

Am Ab Am Ap Ab

+
X Rb Ab X
C1
Cb

1
1
1
Am Ab
d + Rd Q d +
Id Q
Qd +
+
Qd =
X
C1
C1 C2
C1

(2.72)

Decoupler and bell flow variables are related to the velocity of the lumped fluid, by
Qd = Ad X d
Qb = Ab X b

(2.73)

2.5 Mechanical System Model with a Bell

49

where Ad and Ab are the areas of the decoupler and bell orifices. Also, the fluid inertia parameters are related to an equivalent lumped mass, and the resistance is converted to a linear damping
parameter
md
;
A2d
bd
=
;
A2d

Id =
Rd

mb
A2b
bb
Rb = 2
Ab

(2.74)

Ib =

The parameters md and mb represent the mass of the decoupler and bell, respectively. Also, bd
and bb are the linear damping parameters. Next, the compliance variables are transformed from
volumetric stiffness to a linear spring stiffness using,
Cb =

A2m
;
kb

C1 =

A2m
;
k1

C2 =

A2m
k2

(2.75)

where kb , k1 and k2 are the linear spring values representing each volumetric chamber stiffness.
Applying the relations developed in (2.73-2.75) to equation (2.72), gives

Ab
Ab
Ad
Ab
Ab
Ap
k1
1
X
Xb
Xd + kb
1
X
Xb
X =
Am
Am
Am
Am
Am
Am

Am

mb Xb + bb Xb X
Ab

Ab
Ab
Ad
Ad
k1
1
X
Xb
Xd k2
Xd =
(2.76)
Am
Am
Am
Am

Am

md Xd + bd Xd
Ad
To aid the development of a mechanical system that characterizes the equations in (2.76),
the mechanical lever system in Figure 2.19 is considered.

Figure 2.19: Mechanical lever at top of mount.

2.5 Mechanical System Model with a Bell

50

Figure 2.20: Mechanical model of the hydraulic mount with a bell.


The equation for the displacement at Xc is

Ab
Ab
Xc = 1
X
Xb
Am
Am

(2.77)

and the lever is assumed to be weightless, thus the moment balance about X gives
Fc Ab = Fb Am

(2.78)

Applying the lever in Figure 2.19 and equations (2.77) and (2.78) to the moment balance
equations in (2.76), the mechanical system model shown in Figure 2.20 is formed.
To validate the model, the transmitted force to the base is developed and compared to
equation (2.66). Figure 2.21 illustrates the five points of force transmission to the base, which
are calculated as
FT _1 = kr X + br X

Ab
1
X
FT _2 = k1
Am
FT _3 = bd X d

(Am Ab )
FT _4 =
kb
1
Am
Ad
FT _5 = k2
Xd
Am

Ab
Ad
Ad
d
Xb
Xd k2
Xd bd X d md X
Am
Am
Am
Ab
Am

Ab
Ap
Xb
X
Am
Am

(2.79)

2.6 Nonlinear Enhancement of the Bell Model

51

Figure 2.21: Mechanical model indicating transmitted force points.


These equations are transformed to the fluid system form using (2.73) to (2.75) to give:

FT _1 = kr X + br X
FT _2 = (Am Ad )(P1 P2 )
FT _3 = Ad Rd Qd

(2.80)

FT _4 = (Am Ap )Pb
FT _5 = Am P2

Adding the transmitted force terms in (2.80) gives a resultant force term of

FT _1 = kr X + br X + (Am Ad )(P1 P2 ) + Am P2 + Ad Rd Qd (Am Ap )Pb

which is equivalent to (2.66), thus validating the mechanical system model for high frequency
excitations.

2.6 Nonlinear Enhancement of the Bell Model

2.6

52

Nonlinear Enhancement of the Bell Model

The bell system only introduces moderate additions to the nonlinear system equations previously developed. The inertia track and decoupler remain unchanged with respect to nonlinear
modeling. However, with the additional bell system, three chamber compliance values are now
required. New developments are not necessary for both the bell and upper chamber compliances, since equations (2.24) and (2.25) can be used to represent both. Depending on the level
of knowledge of each compliance, either a linear or nonlinear approach can be taken.
The bell flow resistance contributes to the nonlinear mount characteristics and is modeled
with the same approach used for the free decoupler. Starting with the flow resistance for steady
flow through an orifice (2.30), two nonlinear resistance parameters Rb1 and Rb2 are used to
characterize oscillating flow resistance. The nonlinear momentum equation for the bell becomes

+ Rb2 Qb + Ab X (Qb + Ab X)

P1 Pb = Ib Q b + Rb1 (Qb + Ab X)

(2.81)

The nonlinear transmitted force equation follows the same form as equation (2.53), however it receives an additional term representing the bell chamber influence. Using equation (2.66)
the nonlinear transmitted force becomes

FT = kr X + br X + (Am Ad_f nc )(P1 P2 ) + Am P2

(2.82)

+(Ad Ad_holes ) (Rd1 + Rd2 |Qd |) Qd (Am Ap )Pb

where Ad_f nc is the decoupler area switching function (2.52). Equations (2.81) and (2.82) are
used in the linear model presented in (2.61) to (2.66) along with the nonlinear developments in
Section 2.3, to establish the complete hydraulic mount model that includes a bell. This complete
model is assembled in the following summary.

2.7 Summary

2.7

53

Summary

All aspects pertaining to the modeling of the hydraulic mount have been presented in this chapter. A linear model was developed first, and is representative of the cornerstone modeling work
presented by Singh et al. [30]. Bond graphs were used to develop system equations that are
comparable to the results documented in Lee et al. [15] and other publications [5, 3, 18]. Mechanical models developed from the linear systems help to illustrate the operating principles of
hydraulic mounts. These models are similar to the mechanical system representations first developed by Flower [5]. However, this work has mathematically validated the mechanical system
models using the transmitted force.
The development of a nonlinear system model represents advancements to hydraulic mount
modeling. Many nonlinear influences have been accounted for, including the effects of the upper
chamber compliance, inertia track resistance and decoupler switching effects. A summary of the
complete nonlinear system follows. The continuity equations are

C1 P1 = Ap X Qi Qd QL + C1 R1 Ap X Qi Qd
C2 P2 = Qi + Qd + QL

where the volumetric compliance parameters C1 (wdr , VT , Fp ), R1 (wdr , VT , Fp ) and C2 (Fp ) are
dependent on the frequency of excitation wdr , volume amplitude into the upper chamber VT and
the preload force Fp . The momentum equations are
P1 P2 = Ii Q i + (Ri1 + Ri2 |Qi |) Qi

P1 P2 = Id Q d + Rd1 + Rd2 |Qd | + 1 eRd3 Xd arctan(Qd 2 ) Qd


P1 P2 = RL2_f nc |QL | QL
where
RL2_f nc


2
2
= RL2 1 + 7 7 arctan
Vd arctan( 5 Qd ) Vd_ max 6

2.7 Summary

54

Finally, the output transmitted force equation is


FT = kr X + br X + (Ap Ad_f nc )(P1 P2 ) + Ap P2
+(Ad Ad_holes ) (Rd1 + Rd2 |Qd |) Qd
where

Ad_f nc

1
= Ad

arctan
2


2
Vd arctan ( 3 P ) Vd_ max 4

Fp
and the stiffness and damping parameters follow nonlinear behavior captured with kr wdr , X,

Fp ).
and br (wdr , X,

The nonlinear decoupler switching has been approached in several publications [11, 13, 18]
with only moderate success. These nonlinear decoupler models have modeled the switching
conditions with the pressure differential across the decoupler and the volume of fluid through
the decoupler. Although the published models represent the low frequency response, the high
frequency results are incorrect. The nonlinear decoupler model presented in this work operates
using the flow across the decoupler in combination with the volume. This model is effective
over the complete frequency spectrum of interest and will be validated in Chapter 6.
The final portion of this chapter has focused on a model that represents the hydraulic mount
with a bell system. Since, at the time of writing, no published material regarding this type of
system was recognized by the author, this portion of the modeling cannot be compared to other
models.
A lumped parameter model has been established by associating an inertia with the annular column of fluid between the bell plate and mount wall. A mechanical system model was
developed to represent a linear high frequency mount, including the bell and decoupler fluid column inertia. Finally, the nonlinear properties that are imposed by the bell system have been
considered.

2.7 Summary

55

The complete system model that includes the inertia track, decoupler and bell is summarized. The continuity equations are

Cb Pb = Qb (Am Ab Ap )X + Cb Rcb Qb (Am Ab Ap )X

C1 P1 = (Am Ab )X Qb Qd Qi QL
C2 P2 = Qi + Qd + QL

with the nonlinear volumetric parameters Cb (wdr , VT , Fp ), Rcb (wdr , VT , Fp ) and C2 (Fp ). The
bulge damping effect of the new upper chamber is neglected, since the compliance is generally
significantly lower than in the bell chamber.
The momentum equations are
P1 Pb

= Ib Qb + Rb1 (Qb + Ab X) + Rb2 Qb + Ab X (Qb + Ab X)

P1 P2 = Ii Q i + (Ri1 + Ri2 |Qi |) Qi

P1 P2 = Id Q d + Rd1 + Rd2 |Qd | + 1 eRd3 Xd arctan(Qd 2 ) Qd

The leak flow through the closed decoupler is captured using,

P1 P2 = RL2_f nc |QL | QL
where,
RL2_f nc


2
2
= RL2 1 + 7 7 arctan
Vd arctan( 5 Qd ) Vd_ max 6

Finally, the output transmitted force equation is


FT = kr X + br X + (Am Ad_f nc )(P1 P2 ) + Am P2
+(Ad Ad_holes ) (Rd1 + Rd2 |Qd |) Qd (Am Ap )Pb
where,
Ad_f nc

1
= Ad

arctan
2


2
Vd arctan ( 3 P ) Vd_ max 4

2.7 Summary
56

Fp and br wdr , X,
Fp also follow the nonThe stiffness and damping parameters kr wdr , X,
linear behavior captured with

The complete model simulation techniques and validation with experimental system tests
are presented in Chapter 6; while, the following chapter provides a physical explanation of how
these models function.

Chapter

An Insight into the Mathematical Models


This chapter provides physical insight into the hydraulic mount dynamics using the mathematical models developed previously. To begin, the linear system models from Chapter 2 are used
to explain how liquid column resonance influences the output characteristics. The fluid column
resonance is developed in equation form and used to demonstrate the influence of parameters
within the linear model. The physical operation of the nonlinear decoupler model is also investigated using time and frequency domain simulations. Finally, the high frequency linear bell
model is used to explain how the addition of a bell influences the system characteristics. This
linear model is also used to investigate how system parameters influence output.

3.1

Linear Model Characteristics

The linear models developed in Section 2.1.3 are used here to explain how internal liquid column resonance influences mount characteristics.

Using the large amplitude linear model in

equations (2.11) to (2.14),

C1 P1 = Ap X Qi
C2 P2 = Qi
P1 P2 = Ii Q i + Ri Qi
57

3.1 Linear Model Characteristics

58

FT = kr X + br X + Ap (P1 P2 ) + Ap P2
the behavior in the low frequency range is considered.

These equations are manipulated to

create a frequency domain equation for the cross point dynamic stiffness. Using the steady-state
sinusoidal solution detailed in Appendix B equation (B.13), the equation becomes
A2p
C2 A2p j
Ft
K (wdr j) =
= kr + br wdr j +
+
2

C1 Ri C12 C2 wdr + j (Ii C12 C2 wdr


C2 C1 C12 )
X
With the transmitted force FT and input excitation X, the system response is characterized using
the cross-point dynamic stiffness K



FT
K = = |K (wdr j)|

and phase angle

= FT X = ]K (wdr j)

(3.1)

(3.2)

are the steady-state amplitudes of the transmitted force and excitation, respecwhere FT and X
tively.
To physically explain the internal dynamics, the model is simulated using the estimated linear parameters in Table 3.11 . Figure 3.1 shows the dynamic stiffness and phase characteristics
of the model using these estimated parameters. The simulation results and the characteristics
measured and plotted in Figure 1.7a, both indicate the effects of an internal liquid column resonance.
Table 3.1: Linear model parameters.
kr
br
Ap
Ad
C1

=
=
=
=
=

225 N/mm
0.1 103 kg/s
2500 mm2
660 mm2
3.0 104 mm5 /N

C2
Ii
Ri
Id
Rd

=
=
=
=
=

2.6 106 mm5 /N


3.8 106 kg/mm4
10.5 105 kg/s-mm4
7.5 108 kg/mm4
11.7 106 kg/s-mm4

Units selected are those most commonly used in the automotive industry.

59

Phase (Degrees)

Dynamic Stiffness (N/mm)

3.1 Linear Model Characteristics

Frequency (Hz)

Phase (Degrees)

Volume (mm )

Figure 3.1: Dynamic stiffness and phase response of the linear low frequency model.

in-phase

out-of-phase

Frequency (Hz)
Magnitude

Phase

Pressure (Mpa)

Phase (Degrees)

(a)

Frequency (Hz)
Magnitude

Phase
(b)

Figure 3.2: Low frequency linear model response of: (a) volume passing through inertia track;
and (b) pressure in upper chamber.

3.1 Linear Model Characteristics

60

As mentioned in Section 1.2, the liquid column resonance is indicative of the single DOF
(degree-of-freedom) fluid column bound by two compliant chambers. The resonant system is
best illustrated by observing the volume flow through the inertia track, shown in Figure 3.2a. The
volume response resembles a steady-state frequency domain curve of a single DOF system, with
a natural frequency of approximately 15 Hz. Before the system reaches resonance, the inertia
track fluid volume moves relatively in-phase with the excitation. At near-zero frequencies this
volume amplitude is equivalent to the product of excitation amplitude and effective pumping
area. The volume amplitude peaks at a phase of 90 , corresponding to the damped natural
frequency. At frequencies greater than the natural frequency, the volume amplitude diminishes,
moving relatively out-of-phase with respect to the excitation. This effect is sometimes referred
to as flow shut off. The two highlighted regions in Figure 3.2a are used to indicate areas where
significant flow is moving relatively in-phase and out-of-phase with the excitation. Flow through
the inertia track also changes the internal pressure differential, and influences overall hydraulic
mount characteristics.
The upper chamber pressure, illustrated in Figure 3.2b, dominates the hydraulic mount
system response. For frequencies around 10 Hz, the upper chamber pressure amplitude remains
relatively small and out-of-phase. As the pressure increases in magnitude and moves out-ofphase with the excitation, it decreases the dynamic stiffness magnitude and increases the phase
angle. This point is often referred to as the decoupler notch. Between 10 and 15 Hz the
phase angle peaks as increasing pressure begins to push the stiffness back in phase. From 15
to 20 Hz the pressure peaks, since a large amount of fluid is forced into the upper chamber by
the opposing fluid column motion and the excitation pumping motion. As a result, the dynamic
stiffness peaks with a reduced phase angle. Above 20 Hz, the flow magnitude diminishes and
the pressure decreases to a level comparable to a blocked inertia track.

3.1 Linear Model Characteristics

61

An equation for the single DOF system, set up by the inertia and surrounding compliant
chambers, is developed by substituting (2.11) and (2.12) into the derivative of (2.13) to form
i + Ri Q i + (1/C1 + 1/C2 )Qi = Ap X
Ii Q
C1

(3.3)

Using the standard nondimensional form, (3.3) can be written as


i + 2wn Q i + w2 Qi = fo X
Q
n
where the undamped natural frequency of the liquid column is given by
s
(1/C1 + 1/C2 )
wn = 2fn =
Ii

(3.4)

(3.5)

The nondimensional damping ratio is


=

Ri
2Ii wn

(3.6)

Ap
C1 Ii

(3.7)

and the forcing coefficient is


fo =

From equation (3.5), the natural frequency is shown to be dependent only on the upper and
lower chamber compliances and the fluid column inertia. Figure 3.3 illustrates how the natural
frequency corresponds to the peak fluid column motion.
The high frequency linear model characteristics are examined using the equations (2.7)
to (2.10) developed in Section 2.1.3. These equations represent the low amplitude model that
assumes no inertia track
C1 P1 = Ap X Qd
C2 P2 = Qd
P1 P2 = Id Q d + Rd Qd
FT = kr X + br X + (Ap Ad )(P1 P2 ) + Ap P2 + Ad Rd Qd

62

Volume (mm3)

Dynamic Stiffness (N/mm)

3.1 Linear Model Characteristics

Phase (Degrees)

Phase (Degrees)

natural frequency fn

Frequency (Hz)
Stiffness Response

Fluid Column Motion

Volume (mm3)

Dynamic Stiffness (N/mm)

Figure 3.3: Low frequency linear model response, showing the stiffness response and liquid
column motion.

Phase (Degrees)

Phase (Degrees)

natural frequency fn

Frequency (Hz)
Stiffness Response
Fluid Column Motion

Figure 3.4: High frequency linear model response, showing the stiffness response and liquid
column motion.

3.2 Parameter Study for Linear Model

63

The cross point dynamic stiffness is simulated using the values in Table 3.1, and is shown in
Figure 3.4. Since equations (2.7) to (2.10) are of the same form previously examined, the same
internal liquid column resonance behavior results. The high frequency response demonstrates a
liquid column resonance of the decoupler corresponding to a natural frequency of 106 Hz.
From equations (2.7) to (2.9) the natural frequency of the decoupler is determined as:
s
(1/C1 + 1/C2 )
wn = 2fn =
(3.8)
Id

3.2

Parameter Study for Linear Model

Since both linear models have similar dynamic properties, a general model is used to investigate
the influence of parameter changes on the system characteristics.

The general linear model

follows the same form as the two linear models previously investigated, however it is developed
using a generic inertia I and resistance R for the fluid column.

Continuity and momentum

equations form the internal system dynamics as,


C1 P1 = Ap X Q
C2 P2 = Q
P1 P2 = I Q + RQ

(3.9)
(3.10)
(3.11)

and the output, or transmitted force is


FT = kr X + br X + Ap (P1 P2 ) + Ap P2

(3.12)

To simplify the model, this transmitted force is equivalent to equation (2.14), or (2.10) with Ad
set to zero.
Qualitative observations are now made regarding the impact of the compliance parameters, inertia and resistance. Also, the effective pumping area and linear stiffness and damping
parameters are investigated.

64

Phase (Degrees)

Dynamic Stiffness (N/mm)

3.2 Parameter Study for Linear Model

Frequency (Hz)
C1

C1 +20%

C1 +40%

Figure 3.5: Linear model response to changes in the upper chamber compliance parameter.

3.2.1

Compliance

Since the volumetric compliance of the lower chamber is several orders of magnitude higher
than the upper chamber, this parameter has little impact on the overall characteristics. This configuration is referred to as a single-load-bearing chamber mount, or SLBC [6]. If the lower
chamber compliance has a significant magnitude it will influence the dynamic stiffness prior to
resonance and will also act to influence the natural frequency through equation (3.5). In Figure 3.5 the influence of the upper chamber compliance parameter C1 is illustrated. An increase
in the upper chamber compliance decreases the liquid column resonant frequency, according
to equation (3.5). Beyond resonance, the increased compliance decreases the upper chamber
pressure and reduces the dynamic stiffness. The decreased natural frequency will increase the
nondimensional fluid column damping parameter through equation 3.6, thus reducing the peak
phase angle.

65

Phase (Degrees)

Dynamic Stiffness (N/mm)

3.2 Parameter Study for Linear Model

Frequency (Hz)
I

I +20%

I +40%

Figure 3.6: Linear model response to changes in the fluid inertia parameter.

3.2.2

Inertia

The inertia parameter I influences the liquid column resonant frequency, as shown in Figure 3.6.
From equation (3.5), an increase in fluid column inertia will decrease the resulting natural frequency. Also, the additional inertia will increase the amplitude of liquid column motion around
the natural frequency, thereby increasing pressure fluctuations and resulting in higher peaks in
the magnitude and phase characteristics. This point is further reinforced by observing that the
nondimensional damping parameter in equation (3.6) decreases with increased fluid column inertia.

66

Phase (Degrees)

Dynamic Stiffness (N/mm)

3.2 Parameter Study for Linear Model

Frequency (Hz)
R

R +20%

R +40%

Figure 3.7: Linear model response to changes in the fluid resistance parameter.

3.2.3

Resistance

The natural frequency of the liquid column is not altered by the resistance parameter R. Instead,
the resistance is responsible for changing the shape of the liquid column resonance through the
damping equation (3.6). Figure 3.7 illustrates how increased resistance decreases the peaks on
the dynamic stiffness curve.

Physically, the increased resistance limits fluid column motion

during resonance, thus reducing pressure peaks and smoothing the total system response.

67

Phase (Degrees)

Dynamic Stiffness (N/mm)

3.2 Parameter Study for Linear Model

Frequency (Hz)
Ap

Ap +5%

Ap +10%

Figure 3.8: Linear model response for changes in the effective pumping area parameter.

3.2.4

Effective Pumping Area

Figure 3.8 illustrates the impact of the effective pumping area parameter Ap on the system characteristics.

The linear model does not fully capture its significance, since the effective area

parameter is mainly utilized in developing the mounts switching properties. However, it is important to illustrate how the effective area impacts the linear system. As illustrated in Figure 3.8,
and described by equation (3.5) to (3.7), the parameter Ap only influences the driving force on
the fluid column. Increasing the effective area raises the driving function (3.7) and results in
larger fluid column motion. Greater fluid column motion increases internal damping and upper
chamber pressures, resulting in increased peak phase angle and dynamic stiffness.

68

Phase (Degrees)

Dynamic Stiffness (N/mm)

3.3 Nonlinear Decoupler Operation

Frequency (Hz)
kr

kr +20%

kr +40%

Figure 3.9: Linear model response to changes in the linear stiffness parameter.

3.2.5

Linear Stiffness and Damping

The stiffness kr and damping br parameters only influence the characteristics through the transmitted force equation. Since the damping inherent in rubber is relatively small (experimentally
established), the br parameter does not have a significant impact on the system characteristics.
An increase in the real stiffness parameter kr shifts the dynamic stiffness curve as illustrated in
Figure 3.9. Since kr only applies to the real stiffness, a reduction in phase angle results.

3.3

Nonlinear Decoupler Operation

Of the nonlinear developments in Section 2.3, the decoupler switching is the most dominant.
This section examines the model response to confirm functionality and considers the impact of
the main nonlinear decoupler parameter. Frequency domain characteristics are used to explain
the decoupler operation over the complete range of excitations.

3.3 Nonlinear Decoupler Operation

69

To validate the decoupler switching model operation, the linear model is enhanced with
the nonlinear decoupler developments of Section 2.3. These nonlinear equations include the
decoupler momentum, leak flow through the closed decoupler and the transmitted force. All
other nonlinear parameter developments are omitted for this case. The equations become
C1 P 1 = Ap X Qi Qd

(3.13)

C2 P 2 = Qi + Qd

(3.14)

P1 P2 = Ii Q i + Ri Qi

(3.15)

P1 P2 = RL2_f nc |QL | QL

(3.16)

P1 P2 = Id Q d + Rd1 Qd + 1 eRd3 Xd arctan(Qd 2 ) Qd

(3.17)

FT = kr X + br X + (Ap Ad_f nc )(P1 P2 ) + Ap P2 + Ad_f nc Rd1 Qd

(3.18)

where Ad_f nc (2.52) and RL2_f nc (2.58) are dependent on the nonlinear functions of decoupler
area and resistance, respectively. Developed to be


2
2
Vd arctan( 5 Qd ) Vd_ max 6
RL2_f nc = RL2 1 + 7 7 arctan

2
Ad
arctan
Vd arctan ( 3 P ) Vd_ max 4
Ad_f nc =

Due to the nonlinear nature of equations (3.13) to (3.18), a frequency domain equation
form cannot be derived. To obtain frequency response characteristics for this model, the nonlinear equations are simulated in the time domain by solving the set of simultaneous differential
equations, then converted to the frequency domain.
The equations are first rewritten as a set of first order differential equations in the form
y = F (t, y) and solved over a time interval [to , tf ], given initial values yo (to ). The solution
method uses numerical differentiation and a new technique for stiff differential problems [14].
Using MATLAB [19], the set of equations are solved for a particular excitation condition and
a Fourier transform is applied to the steady-state response to achieve the frequency data at the

3.3 Nonlinear Decoupler Operation

70

Table 3.2: Nonlinear model parameters.


kr
br
Ap
Ad
C1
C2

=
=
=
=
=
=

225 N/mm
0.1 103 kg/s
2500 mm2
660 mm2
3.0 104 mm5 /N
2.6 106 mm5 /N

Ri
Ii
Rd
Id
Rd3
RL2

=
=
=
=
=
=

10.5 105 kg/s-mm4


3.8 106 kg/mm4
11.7 106 kg/s-mm4
7.5 108 kg/mm4
31.8 mm1
4.15 106 kg/mm7

maximum amplitude. For each excitation condition the procedure is repeated to determine the
complete frequency response curve.
Further details of the nonlinear simulation technique are covered in Chapter 6, where the
complete nonlinear model is simulated and validated with experimental data.

3.3.1

Low Frequency Response

The decoupler switching characteristics are investigated for low frequency large, amplitude excitations by simulating the model with parameters given in Table 3.2. Excitations having peakto-peak amplitude levels of 1 mm and 2 mm are applied at 6 Hz to observe the decoupler volume, inertia track volume and upper chamber pressure. As Figure 3.10 indicates, the decoupler
equations impose the desired nonlinear mount characteristics. At approximately 350 mm3 the
decoupler plate contacts the cage, stopping flow and pushing fluid through the inertia track, as
illustrated in Figure 3.10b. These time domain plots also indicate the amplitude sensitivity inherent in the model. As the excitation amplitude decreases, less fluid is pushed into the upper
chamber, which reduces decoupler cage contact time. Since the decoupler is free a greater portion
of time, the upper chamber pressure and inertia track motion are both reduced. This characteristic is also observed in the frequency domain system response of Figure 3.11; dynamic stiffness is
reduced after resonance due to the reduction in upper chamber pressure. General characteristics
of the nonlinear system under large amplitude low frequency excitations are also comparable to
the linear model with only an inertia track.

71

Decoupler
Volume (mm3)

3.3 Nonlinear Decoupler Operation

Inertia Track
3
Volume (mm )

(a)

Upper Chamber
Pressure (MPa)

(b)

Time (s)
(c)
2mm P-P

1mm P-P

Figure 3.10: Time domain simulation of nonlinear decoupler model: (a) decoupler volume; (b)
inertia track volume; (c) upper chamber pressure.

72

Phase (Degrees)

Dynamic Stiffness (N/mm)

3.3 Nonlinear Decoupler Operation

Frequency (Hz)
2 mm P-P

1 mm P-P

Figure 3.11: Low frequency simulation of nonlinear model.

3.3.2

High Frequency Response

The same continuous nonlinear model is now simulated with the parameters in Table 3.2 over
0 to 250 Hz, illustrated in Figure 3.12. For an excitation amplitude of 0.05 mm the nonlinear decoupler model demonstrates characteristics corresponding to the linear model response in
Figure 3.4. Excitation levels of 0.1 mm and 0.2 mm also demonstrate the amplitude-sensitive
characteristics in the high frequency range.

With increased excitation amplitude, the decou-

pler fluid column motion enters the nonlinear region around resonance, shown in Figure 3.13.
Since the decoupler accommodates approximately 300 mm3 , the volume of fluid displaced into
the upper chamber is limited and the peak upper chamber pressure is diminished. These effects
influence the system characteristics by reducing the peak dynamic stiffness and phase angle.
The frequency response characteristics in Figures 3.12 and 3.13 introduce an interesting
phenomena. Since the new hydraulic mount model is now a nonlinear set of differential equations, the frequency response is multi-valued. The resulting steady-state solution now has more

73

Phase (Degrees)

Dynamic Stiffness (N/mm)

3.3 Nonlinear Decoupler Operation

Frequency (Hz)
0.05 mm P-P

0.075 mm P-P

0.1 mm P-P

Phase (Degrees)

Volume
3
Magnitude (mm )

Figure 3.12: Dynamic stiffness characteristics of the nonlinear model under high frequency
excitations.

Frequency (Hz)
0.05 mm P-P

0.075 mm P-P

0.1 mm P-P

Figure 3.13: Decoupler motion in the nonlinear model under high frequency excitations.

3.3 Nonlinear Decoupler Operation

74

than one possible value for a given excitation condition. This system response is not only dependent on the amplitude and frequency of excitation, but is also dependent on the initial conditions.
As cited in Inman [7], a unique characteristic of nonlinear systems is a bend in the frequency
response, produced by unstable regions. To demonstrate that this unstable region exists within
the nonlinear decoupler model, frequency response curves have been calculated by sweeping
the excitations. Using this technique, the initial conditions for a particular frequency become
the last conditions of the previous response. The impact of the initial conditions can be illustrated by comparing the frequency response for forward swept excitations and backward swept
excitations.
Figure 3.14 shows how the amplitude and phase of the decoupler motion are dependent
on excitation sweep direction, furthermore the unstable region becomes evident. However, the
impact of this phenomenon is somewhat diluted in the system dynamic characteristics. As Figure 3.15 illustrates, the dynamic stiffness is not greatly influenced by initial conditions, however,
the phase does show a small region of instability.
To compare this simulation approach with the technique used for experimental evaluation,
both Figures 3.15 and 3.14 include the frequency response obtained using a mean dwell method.
This method brings the system to rest for a period of time between excitations, to allow all
internal dynamics to settle and resetting all initial conditions to zero. From the results, it is
concluded that an acceptable frequency response is obtained by setting the initial conditions to
zero for each excitation condition. Since the mount testing procedure also uses a mean dwell sine
sweep process, all further nonlinear simulations will be conducted by setting initial conditions to
zero for each excitation.

75

Phase (Degrees)

Volume
Magnitude (mm3)

3.3 Nonlinear Decoupler Operation

Frequency (Hz)
Forward Sweep

Backward Sweep

Mean Dwell

Phase (Degrees)

Dynamic Stiffness (N/mm)

Figure 3.14: Decoupler motion indicating region on instability and dependence on sine sweep
conditions.

Frequency (Hz)
Forward Sweep

Backward Sweep

Mean Dwell

Figure 3.15: Dynamic stiffness and phase indicating impact of sine sweep conditions.

3.4 Linear Bell Model Characteristics

3.3.3

76

Influence of Nonlinear Decoupler Parameters

The switching characteristics of the nonlinear decoupler model are controlled with the parameter
Rd3 . This parameter limits the fluid volume through the decoupler via equation (3.17), reflecting the decoupler cage height. To verify the perceived impact of the decoupler parameter Rd3
on the system response, the model is simulated with an excitation of 6 Hz and 2 mm P-P. Using the parameters in Table 3.2, the time domain decoupler volume Vd is plotted in Figure 3.16.
Simulation results verify a direct relation between Rd3 and the volume of fluid passing through
the decoupler. Increasing the decoupler parameter limits the volume of fluid through the decoupler, indicative of a decreased cage height. Details of this parameter-to-geometry relation are
presented in Chapter 7.
To complete the discussion on the operation of the nonlinear decoupler model, the parameter Vd_ max is considered. This parameter represents the maximum fluid volume amplitude
allowed through the decoupler. Since this characteristic is also related to the decoupler cage
height it will be shown that Vd_ max is a function of the decoupler parameter Rd3 . The relation-

Decoupler
Volume (mm3)

ship between these two parameters is also presented Chapter 7.

Time (s)
Rd3

Rd3 +30%

Figure 3.16: Simulation of decoupler volume indicating the response to changes in the decoupler parameter.

3.4 Linear Bell Model Characteristics

3.4

77

Linear Bell Model Characteristics

A bell plate is incorporated into the hydraulic mount to alleviate the undesired high frequency
resonance of the decoupler liquid column. The bell model, introduced in Section 2.4, assigns
inertia and flow parameters to the annular column of fluid between the bell plate and mount wall.
This new bell fluid column places an additional DOF within the mount. Experimental studies
show that the bell is only effective at high frequencies. This can be explained by noticing that
the bell inertia is several orders of magnitude less than that of the inertia track. Thus, the low
frequency high amplitude characteristics are not significantly altered. This section, therefore,
only examines the high frequency response of the linear bell and decoupler model.
The high frequency linear model that includes the decoupler and bell dynamics (with Qi
set to zero), is developed using equations (2.60) to (2.66). The linear equations become
Cb P b = Qb (Am Ab Ap )X

(3.19)

C1 P1 = (Am Ab )X Qb Qd

(3.20)

C2 P2 = Qd

(3.21)

P1 Pb = Ib Q b + Rb (Qb + Ab X)

(3.22)

P1 P2 = Id Q d + Rd Qd

(3.23)

FT = kr X + br X + (Am Ad )(P1 P2 ) + Am P2 + Ad Rd Qd (Am Ap )Pb

(3.24)

The two degrees of freedom now present within the system, stem from the momentum
equations (3.22) and (3.23). To simulate the frequency response of this system, equations (3.19)
to (3.23) are first manipulated into the matrix form in Appendix B equation (B.21)

Ib 0
0 Id

b
Q
d
Q

Rb 0
0 Rd

Q b
Q d
"

Am Ab
C1

1
C1

+
1
C1

1
Cb

1
C1

Am Ap Ab
Cb
Am Ab
C1

1
C1

+
#

1
C2

Qb
Qd

Rb Ab
0

3.4 Linear Bell Model Characteristics

78

This system of equations is then converted to the frequency domain assuming the steady-state
solutions
b ejwdr t
Qb = Q
d ejwdr t
Qd = Q

(3.25)

jwdr t
X = Xe
The frequency domain responses of Qd and Qb are obtained by solving the characteristic equation and isolating the variables in matrix form. With the flow equations, the pressure
response in each chamber is developed and used to calculate the transmitted force via equation (3.24). The cross point dynamic stiffness K and phase angle , are determined using the
same approach presented in Section 3.1. Appendix B contains details on the equation manipulation for simulation.
Table 3.3: Parameters used in linear bell model.
kr
br
Ap
Ab
Ad
Am

=
=
=
=
=
=

225 N/mm
0.1 103 kg/s
2500 mm2
600 mm2
660 mm2
4900 mm2

Cb
C1
C2
Rd
Id
Rb
Ib

=
=
=
=
=
=
=

3.0 104 mm5 /N


700 mm5 /N
2.6 106 mm5 /N
11.7 106 kg/s-mm4
7.5 108 kg/mm4
2.8 107 kg/s-mm4
2.0 108 kg/mm4

Using estimated parameters, given in Table 3.3, the high frequency dynamic stiffness and
phase responses are simulated and shown in Figure 3.17. Response characteristics illustrate how
the additional bell plate alleviates the increased stiffness from decoupler resonance. To explain
physically the internal dynamics generating these unique characteristics, the frequency response
of both bell and decoupler fluid columns are considered along with the chamber pressures in
Figures 3.18 and 3.19.
The two coupled differential equations in (B.21) represent a two degree of freedom system that contains eigenvalues, or natural frequencies, solved through the characteristic equation.

79

Phase (Degrees)

Dynamic Stiffness (N/mm)

3.4 Linear Bell Model Characteristics

Frequency (Hz)

Figure 3.17: Dynamic stiffness and phase characteristics from the simulated linear model that
includes the bell and decoupler.
Each natural frequency corresponds to mode shapes pertaining to the internal fluid column variables Qb and Qd , flow through the bell and decoupler respectively. Following the same approach
as in Section 3.1, the fluid column motions are correlated to the internal pressures which directly
influence the systems dynamic stiffness characteristics.

Since the system now exhibits two

modes of vibration, frequency response curves indicate characteristics not present in the standard hydraulic mount configuration.
Table 3.4: Frequency and mode shape vectors of high frequency bell model.

Natural Frequency:

Mode 1
94 Hz

Mode 2
1525 Hz

Mode Shape:

(1, 1.02)T

(1, 0.26)T

From the eigenvalue solution in Table 3.4, the system undergoes a first resonant frequency
at 94 Hz. For illustrative purposes the phase in Figure 3.18 has been adjusted 180 to be aligned
with respect to the excitation; the original coordinate direction of Qb is opposite that of Qd .

80

Volume (mm3)

Phase (Degrees)

3.4 Linear Bell Model Characteristics

Phase

Frequency (Hz)
Magnitude
(b)

Phase

Volume (mm3)

Phase (Degrees)

Frequency (Hz)
Magnitude
(a)

Pressure (Mpa)

Phase (Degrees)

Figure 3.18: High frequency linear model response of: (a) decoupler fluid column motion; and
(b) bell fluid column motion.

Phase

Frequency (Hz)
Magnitude
(b)

Phase

Pressure (Mpa)

Phase (Degrees)

Frequency (Hz)
Magnitude
(a)

Figure 3.19: High frequency linear model response of: (a) upper chamber pressure; and (b)
bell chamber pressure.

3.4 Linear Bell Model Characteristics

81

During this mode of excitation the decoupler and bell oscillate at 90 phase to the excitation.
Physically, this mode of vibration represents both decoupler and bell inertia values moving inphase, thus the first natural frequency has decreased to 94 Hz, from the previous value of 106 Hz
with just a decoupler. This trend follows from equation (3.8), where increases in inertia decrease
the natural frequency. The resulting dynamic stiffness and phase characteristics represent those
of a standard hydraulic mount up to approximately 100 Hz.
At frequencies between 100 and 150 Hz, decoupler motion decreases drastically at a phase
of 180 to the excitation. Pressure in the upper chamber also decreases rapidly in this frequency range, as the bell fluid column motion moves 180 to the excitation and accommodates
fluid displaced into the upper chamber.

As upper chamber pressure decreases, the dynamic

stiffness also undergoes the same trend, producing another low point or notch in the curve in
Figure 3.17. From this explanation, it will be shown in Chapter 6 that even without a decoupler
the mount exhibits the same decrease in stiffness.
The frequency response of the decoupler volume indicates an anti-resonance occurring at
156 Hz (see Figure 3.18a). Antiresonance is a local effect in MDOF systems, occurring only
in certain response curves depending on the modal coordinates. This is explained further in Maia
et al. [17]. Physically, the antiresonance implies the fluid column is grounded at this frequency
or its real motion is zero. Also characteristic of the antiresonance is the 180 phase shift, which
is shown in Figure 3.18a.
At frequencies beyond 156 Hz the decoupler motion begins to increase, now in-phase with
the excitation. The upper chamber pressure amplitude also increases at 180 to the excitation.
As the pressure increases it drives the dynamic stiffness up and forces the response out of phase.
This trend continues through the frequency range of interest.
From Table 3.4, the second resonant frequency in the system occurs at 1525 Hz and has
a mode shape with the decoupler and bell fluid column 180 out of phase. This fluid collision

3.5 Parameter Studies on a Mount with the Bell System

82

in the upper chamber results in high pressures and high dynamic stiffness. The design of a hydraulic mount with a bell system will, therefore, not remove the internal liquid column resonance
of the decoupler, but will give the engineer greater ability to tune high frequency characteristics.

3.5

Parameter Studies on a Mount with the Bell System

The additional bell dynamics create a system capable of achieving better isolation characteristics
in the high frequency range. However, the multiple degrees of freedom within the system also increase the complexity. Since the characteristic equation for the system must be solved to obtain
the eigenvalues, the natural frequencies cannot be written explicitly in terms of the parameters.
The natural frequencies and mode shape vectors in Table 3.4 are approximated using the mass
and stiffness matrices from equation (B.21) and solving the eigenvalue problem, given known
parameters. To better understand how the parameters influence the system response, the antiresonance of the decoupler will be considered along with simulations illustrating the influence of
parameter changes.
As previously discussed, the antiresonance frequency of the decoupler fluid column is associated with the notch point in the dynamic stiffness curve. Theoretically, when the equation
representing the decoupler frequency response in Figure 3.18a is zero, the decoupler fluid column is at antiresonance. From equations (3.19) to (3.23), the frequency domain equation for the
decoupler flow is determined in Appendix B equation (B.27), as

d =
Q

(Am Ab )wdr
C1

1
Cb

Am Ap Ab
(Am Ab )Cb

2
Ib wdr

+ Rb

Characteristic Equation

Rb Ab
Am Ab

wdr j

where the Characteristic Equation is the determinant of the inertia, resistance and compliance
matrix transformed into the frequency domain.

Since the decoupler fluid column motion is

zero at antiresonance, it also corresponds to the point of zero flow. Therefore, the frequency
of antiresonance occurs when the real component of the numerator in equation (B.27) is zero.

3.5 Parameter Studies on a Mount with the Bell System


Solving for the antiresonance driving frequency wdr yields
v

u
u 1 1 Am Ap Ab
t Cb
Am Ab
wdr = wa = 2fa =
Ib

83

(3.26)

where wa and fa denote the notch or antiresonance frequency in radians per second and Hertz,
respectively. Substituting the parameters from Table 3.3 into equation (3.26) yields a frequency
of 156 Hz, corresponding to the notch frequency previously observed.

Equation (3.26) also

highlights that antiresonance is not a function of decoupler inertia or middle chamber compliance, but rather a function of bell chamber compliance, fluid column inertia and effective area
parameters within the mount.
The influence of equation (3.26) will be confirmed by observing the effect on model simulation results to changes in compliance, inertia, resistance and area parameters. Stiffness and
damping parameters kr and br have been omitted from this analysis since they were shown previously to have minimal impact on the system response.

3.5.1

Compliance

Two of the three compliance values will be investigated. The lower chamber compliance is omitted since it has an insignificant impact on the system characteristics, leaving the bell and upper
chamber to be studied. Figure 3.20 illustrates changes in system characteristics to variations in
the upper chamber compliance C1 parameter. Since the compliance of this chamber is several
orders of magnitude smaller than the bell compliance, these simulations require a large percentage increase in C1 to illustrate the influence. From the calculated resonant frequencies tabulated
in Table 3.5 an increase in the upper chamber compliance will reduce both natural frequencies,
however the second mode indicates a larger decrease. Dynamic stiffness and phase responses
indicate the forward shift in liquid column resonances and a stationary antiresonance frequency,
corresponding to equation (3.26). Both the tabulated data and the frequency response confirm
that the notch location is not influenced by the upper chamber compliance.

3.5 Parameter Studies on a Mount with the Bell System

84

Table 3.5: Frequency modifications to changes in the upper chamber compliance parameter.
Parameter
C1
C1 +1000%
C1 +3000%

Mode 1
94 Hz
88 Hz
78 Hz

Mode 2
1525 Hz
514 Hz
336 Hz

Antiresonance
156 Hz
156 Hz
156 Hz

The bell chamber compliance parameter Cb is also considered. As Table 3.6 indicates, the
bell compliance influences the first resonant frequency, while the high frequency mode remains
relatively unchanged. Increases in the bell compliance parameter decrease the first mode of vibration and shifts the notch frequency, shown in Figure 3.21. Changes in antiresonance location
are again indicated by equation (3.26). Note that changes in system characteristics at the first
resonance are similar to results found in the single DOF system (see Figure 3.5).
Table 3.6: Frequency modifications to changes in the bell chamber compliance parameter.
Parameter
Cb
Cb +40%
Cb +80%

Mode 1
94 Hz
80 Hz
71 Hz

Mode 2
1525 Hz
1521 Hz
1520 Hz

Antiresonance
156 Hz
132 Hz
116 Hz

85

Phase (Degrees)

Dynamic Stiffness (N/mm)

3.5 Parameter Studies on a Mount with the Bell System

Frequency (Hz)
C1

C1 +1000%

C1 +3000%

Phase (Degrees)

Dynamic Stiffness (N/mm)

Figure 3.20: Dynamic stiffness and phase response to changes in the upper chamber compliance parameter.

Frequency (Hz)
Cb

Cb +40%

Cb +80%

Figure 3.21: Dynamic stiffness and phase response to changes in the bell chamber compliance
parameter.

3.5 Parameter Studies on a Mount with the Bell System

3.5.2

86

Inertia

Figure 3.22 and Table 3.7 highlight the impact of the decoupler fluid column inertia parameter Id
on the system characteristics. An increase in decoupler inertia decreases both natural frequencies within the system. Figure 3.22 demonstrates that dynamic stiffness characteristics at the
first resonance change shape similarly to the single inertia system illustrated in Figure 3.6. Furthermore, antiresonance within the system is not influenced by changes to the decoupler inertia.
Table 3.7: Frequency modifications resulting from changes in the decoupler fluid column inertia parameter.
Parameter
Id
Id +40%
Id +80%

Mode 1
94 Hz
62 Hz
74 Hz

Mode 2
1525 Hz
1480 Hz
1454 Hz

Antiresonance
156 Hz
156 Hz
156 Hz

The impact of the bell fluid column inertia Ib is illustrated in Figure 3.23 and Table 3.8.
The plot shows little change in the first natural frequency to increases in the bell column inertia, however the second mode frequency decreases significantly. Also, the response shows a decreased antiresonance frequency as the notch shifts to the right, corresponding to equation (3.26).
Since the bell fluid column inertia is significantly lower than that of the decoupler, the first mode
of vibration does not undergo much change as the bell inertia is altered.
Table 3.8: Frequency modifications resulting from changes in the bell fluid column inertia
parameter.
Parameter
Ib
Ib +40%
Ib +80%

Mode 1
94 Hz
91 Hz
87 Hz

Mode 2
1525 Hz
1341 Hz
1226 Hz

Antiresonance
156 Hz
132 Hz
117 Hz

87

Phase (Degrees)

Dynamic Stiffness (N/mm)

3.5 Parameter Studies on a Mount with the Bell System

Frequency (Hz)
Id

Id +40%

Id +80%

Phase (Degrees)

Dynamic Stiffness (N/mm)

Figure 3.22: Dynamic stiffness and phase response to changes in the decoupler fluid column
inertia parameter.

Frequency (Hz)
Ib

Ib +40%

Ib +80%

Figure 3.23: Dynamic stiffness and phase response to changes in the bell fluid column inertia
parameter.

88

Phase (Degrees)

Dynamic Stiffness (N/mm)

3.5 Parameter Studies on a Mount with the Bell System

Frequency (Hz)
Rd

Rd +40%

Rd +80%

Figure 3.24: Dynamic stiffness and phase response to changes in the decoupler resistance
parameter.

3.5.3

Resistance

Figure 3.24 includes the frequency response curves of dynamic stiffness and shows the impact of
changes in the decoupler resistance parameter Rd . Since the resistance is independent of internal
stiffness and damping, changes in resistance have no influence on the resonance or antiresonance
frequencies within the system. Instead, decoupler resistance is shown to decrease the peaks on
the first mode, similar to changes in the single DOF system illustrated in Figure 3.7.
Changes to the bell resistance parameter Rb have no significant impact on the system characteristics. The bell resistance parameter reflects the resistance to fluid flow between the mount
wall and the bell plate, and generally has values that are several orders of magnitude less than
the decoupler resistance.

89

Phase (Degrees)

Dynamic Stiffness (N/mm)

3.5 Parameter Studies on a Mount with the Bell System

Frequency (Hz)
Am

Am +10%

Am +20%

Figure 3.25: Dynamic stiffness and phase response to changes in the mount area parameter.

3.5.4

Area Parameters

Finally, the influence of internal area parameters are considered. First it should be highlighted
that the natural frequencies for both modes of vibration are not influenced by any area parameters,
since the eigenvalue problem only uses the inertia and compliance matrices of equation (B.21).
Table 3.9 and Figure 3.25 provide evidence that Am primarily changes the antiresonance frequency, or notch of the dynamic stiffness characteristics. These results are again supported by
equation (3.26).
Table 3.9: Frequency modifications resulting from changes in the mount area parameter.
Parameter
Am
Am +10%
Am +20%

Mode 1
94 Hz
94 Hz
94 Hz

Mode 2
1525 Hz
1525 Hz
1525 Hz

Antiresonance
156 Hz
148 Hz
141 Hz

3.5 Parameter Studies on a Mount with the Bell System

90

The effective pumping area Ap also influences the notch frequency. Figure 3.26 shows how
the changes in stiffness response at the first resonance are similar to the single DOF system in
Figure 3.8. As the effective area increases, more fluid is pushed through the system, increasing
damping and pressure levels during resonance. The antiresonance or notch frequency is also
altered through equation (3.26) and is tabulated in Table 3.10.
Table 3.10: Frequency modifications resulting from changes in the effective pumping area
parameter.
Parameter
Ap
Ap +10%
Ap +20%

Mode 1
94 Hz
94 Hz
94 Hz

Mode 2
1525 Hz
1525 Hz
1525 Hz

Antiresonance
156 Hz
164 Hz
171 Hz

Subtle changes in the notch frequency result from changes in the bell area Ab . Figure 3.27
gives an indication of simulation results for increased bell area parameter. Table 3.11 reinforces
that the antiresonance increases with increases in Ab .
Table 3.11: Frequency modifications resulting from changes in bell area parameter.
Parameter
Ab
Ab +40%
Ab +80%

Mode 1
94 Hz
94 Hz
94 Hz

Mode 2
1525 Hz
1525 Hz
1525 Hz

Antiresonance
156 Hz
161 Hz
166 Hz

A cautionary note should be made about the Am and Ab parameters. In most cases, modifying these parameters will alter the bell inertia and resistance parameters, yielding slightly different results than shown here. Chapter 7 will present some preliminary suggestions and results
on the relation between Am , Ab and the bell fluid column parameters.

91

Phase (Degrees)

Dynamic Stiffness (N/mm)

3.5 Parameter Studies on a Mount with the Bell System

Frequency (Hz)
Ap

Ap +10%

Ap +20%

Phase (Degrees)

Dynamic Stiffness (N/mm)

Figure 3.26: Dynamic stiffness and phase response to changes in the effective pumping area.

Frequency (Hz)
Ab

Ab +40%

Ab +80%

Figure 3.27: Dynamic stiffness and phase response to changes in the bell area parameter.

3.6 Summary

3.6

92

Summary

The main focus of this chapter was to provide a better understanding of the models developed
in Chapter 2.

By investigating the simulation results, it was first concluded that the models

demonstrate similar trends to experimental data. After the linear models were deemed feasible,
they were used to physically explain the dynamics within the mount. Numerical studies were
conducted using the linear models to highlight the impact of system parameters on the overall
characteristics.
Preliminary nonlinear model simulations showed how this model captures the desired response across the complete frequency and amplitude spectrum of interest. Insight into the decoupler switching parameters was also documented.
Finally, the bell model has been shown to contain additional properties for tuning high
frequency characteristics within the mount. Dynamic characteristics of the bell and decoupler
system were physically explained using a linear, high frequency model. By applying modal
analysis techniques the high frequency notch, indicative of a bell system, has been associated
with an antiresonance of the decoupler fluid column. After physically explaining the system
response, the influence of parameters was investigated.
All work conducted in this chapter gives a reasonable understanding of the mount operation, however there is still a lingering question. What are the actual parameter values within the
system that pertain to a particular hydraulic mount? The next half of this thesis is dedicated to
resolving this issue.

Chapter

Experimental Design
Up to this point, the hydraulic mount has been modeled and simulated to demonstrate the internal
dynamics and influence of system parameters. The thesis now moves into quantifying these
system parameters by introducing an experimental apparatus designed to isolate and identify the
hydraulic mount system parameters.
The first portion of this chapter is dedicated to establishing the goals and approach to the
experimental process. The available resources and identification techniques are developed into
experimental design criteria, which are then used for the final design.

Instrumentation and

computer interfacing is discussed and calibration procedures are presented.

4.1

Goals and Approach

One of the primary purposes of this research is to accelerate hydraulic mount design. System
models and simulations presented thus far, give a good indication of the function and influence
of several parameters; however, the models lack some relation to actual hydraulic mount components. As a result, this work begins to focus on measuring parameters from existing hydraulic
mount components.
The approach is conceptually simple. Since all models are based on lumped parameter
assumptions, it is possible to break apart the mount and identify parameters in isolation. For

93

4.2 Design Criteria

94

example the inertia track momentum equation (2.29)

P1 P2 = Ii Q i + (Ri1 + Ri2 |Qi |) Qi


completely characterizes this component through the parameters Ii , Ri1 and Ri2 . If the inertia
track is isolated and a controlled flow input is applied, the pressure response from this equation
can be used to determine the unknown parameters.
Component isolation and parameter identification is the first step to bridging the gap between the mathematical models and physical hydraulic mount systems (see Figure 4.1).
- Isolate components
- Identify parameters

Nonlinear Lumped
Parameter Models

Physical System
Response

Figure 4.1: Experimental Approach

4.2

Design Criteria

The experimental apparatus was designed to satisfy several criteria.

From the identification

perspective, the experimental system had to be capable of providing oscillatory input conditions
to match those found within the mount. Also, the data acquisition system had to be capable of
reading small deviations in pressure and transient flow in the time domain. Available equipment
limitations required that the flow be selected as the controlled input to the apparatus. The system
had to operate with internal pressures of 0 to 50 psi (gauge) with the capability to safely reach
100 psi. These criteria were established from hydraulic mount pressure measurements cited in
Kim [10].

4.3 Final Design

95

When establishing design criteria it was determined that an existing MTS servo-controlled
hydraulic rate machine (mount testing system) could provide high frequency controlled excitation to the apparatus. This approach reduced the cost and complexity of the system while
utilizing high precision, computer interfaced equipment.
The physical construction of the apparatus had to accommodate all components while interfacing both mechanically and electrically with the rate machine. Attention was also placed
on a rigid apparatus design to reduce any undesired dynamics during testing.

4.3

Final Design

The experimental apparatus in Figure 4.2 (photographs in Figure 4.3 and 4.4) uses a hydraulic
cylinder
1 , driven via the servo-controlled actuator. With the base of the cylinder clamped
to the machine base, linear motion of the actuator displaces fluid into a two chamber vessel.
Displacement control on the actuator becomes volume control by multiplying the fixed piston
area by the linear actuator motion. While keeping the piston top open to atmosphere, actuator
2 into the identification vessel
3.
motion pushes fluid through a 12 -in diameter union joint
The 5 12 -in inside diameter vessel is constructed with 12 -in thick cast acrylic, contained between
aluminum connectors.

The modular design uses holding plates


4 , clamped between vessel

sections, to fix the various components in line with the controlled flow. High frequency pressure
sensors
5 are used to collect output data on each side of the holding plate, capturing the pressure
differential. Complete assembly drawings and a material list of the final design are compiled in
Appendix C. Details on the setup of specific component tests is presented in Chapter 5.

4.3 Final Design

96

Excitations from
servo-controlled actuator

5
4

Base of MTS Rate Machine

Figure 4.2: Schematic of experimental test apparatus.

4.3 Final Design

Figure 4.3: Photo of experimental apparatus in MTS servo-controlled rate machine.

97

4.3 Final Design

Figure 4.4: Photograph of experimental apparatus.

98

4.4 Instrumentation

Servo-controlled
Actuator

Experimental
Test Rig
Pressure
Transducers

99

LVDT

Controller
Command
A/C Conditioner
D/C Conditioners
Data Acquisition

Workstation

Figure 4.5: Instrumentation setup for the experimental apparatus.

4.4

Instrumentation

The apparatus design made full use of the existing equipment so that the only new instrumentation required were two pressure transducers. Two Sensotec subminiature Model S pressure
transducers were selected for their high natural frequency, enabling accurate readings for transient pressure measurements. Since pressure and force transducers are electrically similar, the
two pressure sensors were interfaced directly to the MTS controller through existing force transducer ports and internal signal conditioners. The pressure transducers were directly conditioned
and sampled using the existing real-time high speed, data acquisition system. In some instances
the pressure signals were even used for feedback on the actuator. This arrangement enabled
precise pressure measurements to be acquired with actuator position readings, necessary for
measuring transient behavior. Figure 4.5 illustrates the instrumentation configuration.

4.5 Calibration

4.5

100

Calibration

To interface the pressure transducers with the MTS controller software, a sensor calibration file
was created using the transducer calibration procedure [21]. Pressure sensor calibration was
conducted using a certified measurement tool, consisting of a pneumatic pump and digital pressure gauge, illustrated in Figure 4.6. Pressure signals were calibrated for 0 to 50 psi and the
units were left in Imperial units until post data collection processing.
Pressure
Transducer

air lines
Signal to controller
Pneumatic
Hand Pump

Digital Pressure
Guage

Figure 4.6: Illustration of pressure sensor calibration tool.

Curiously, the pressure signals were calibrated such that a positive pressure within the
vessel reads as negative. Therefore, the maximum pressure of 50 psi was actually read as 50 psi
by the MTS system and in the data. This configuration used the same convention for the existing
actuator calibration, measuring a downward motion as negative, and was necessary for pressure
feedback control within the vessel.
In addition, the 0 to 50 psi calibrated pressure range does not accommodate possible vacuum conditions within the vessel. To ensure that locations within the test apparatus never droped
below atmospheric pressure, all tests were conducted under elevated pressure conditions. Details
of the mean operating pressure will be stated as the individual tests are documented.

4.5 Calibration

4.5.1

101

Vessel Chamber Compliance

The actuator displacement measurement Xact from the LVDT are multiplied by the hydraulic
piston area Apiston to establish the volume of fluid that passes into the vessel, or volume excitation
Vexcit ,
Vexcit = Xact Apiston

(4.1)

where the piston area is 5350 mm2 . The flow is then calculated as the time derivative of the
volume excitation
Qexcit = V excit

(4.2)

Although the test apparatus was designed to be as rigid as possible, a portion of fluid from
the excitation volume Vexcit will be accommodated by the volumetric expansion of the vessel. To
determine the actual fluid volume that passes through the holding plate Vactual it is first necessary
to identify the compliance associated with the lower vessel Crig , as illustrated in Figure 4.7. The
damping associated with this expansion is assumed negligible.

Xact

Plate used to find


compliance Crig

Ps2

Vactual
Apiston

Ps1
Crig

Vexcit

Figure 4.7: Illustration of the volumetric compliance associated with the vessel lower chamber.

To determine the volumetric compliance, the flow across the holding plate is first blocked
with a rigid plate. Then, using the pressure sensor for feedback, the lower vessel chamber is

4.5 Calibration

102

Lower Chamber Pressure (psi)


-40

-35

-30

-25

-20

-15

-10

-5

-0.1
-0.2
-0.3
-0.4

Actuator Displacement (mm)

-0.5
Measurements

Trendline

Figure 4.8: Plot of lower chamber pressure versus actuator displacement.


pressurized using the actuator driven piston. Data from the pressure sensor Ps1 and actuator
displacement Xact are plotted in Figure 4.8.

An equation representing the measured data is

established from the trendline to be,


2
Xact = 1.2 104 Ps1
+ 0.0175Ps1

(4.3)

Multiplying by the piston area (4.1) gives the volume displaced in the lower chamber. Using this
volume, the slope at some instantaneous pressure Pinst in the lower vessel chamber represents
the compliance
Crig =

Apiston Xact
= Apiston 2.4 104 Pinst + 0.0175
Ps1

(4.4)

To determine the actual volume of fluid passing through the holding plate during normal
operation, the time domain measurements of pressure Ps1 are used to modify actuator displacement Xact . First, the differential data series Xact and Ps1 are calculated using,
k+1
k
Xact = Xact
Xact

Ps1 = Ps1 k+1 Ps1 k

(4.5)

4.6 Apparatus Capabilities

103

where the change in excitation volume is


Vexcit = Apiston Xact

(4.6)

The instantaneous pressure in the lower chamber is then approximated as the average between
the k and k + 1 measurements
k+1

k
Pinst = Ps1
+ Ps1
/2

(4.7)

Using equation (4.4) the lower chamber compliance is then calculated and used with Ps1 to
modify the volume displaced by the actuator. The result is the actual volume change, written as
Vactual = Vexcit Crig Ps1

(4.8)

With the actual change in volume passing through the holding plate (4.8), the cumulative volume
Vactual is determined using
k+1
k
Vactual
= Vactual
+ Vactual

(4.9)

From the preceding development, the volumetric compensation procedure can only be conducted in the time domain. To obtain frequency domain measurements, the actuator position and
chamber pressures data must be first collected in the time domain. At the end of each frequency
and amplitude test condition, both input excitation and output pressures are sampled and written
to a file. The time domain data is processed using equations (4.1) to (4.9) to calculate the actual
volume, or flow, passing into the system. Frequency analysis is then conducted on the output
pressures and calculated input. For all analysis procedures presented in the following chapter,
the input to the isolated system will be compensated to account for the vessel compliance.

4.6

Apparatus Capabilities

The experimental apparatus has safety and accuracy limitations.

For reliable operation, the

pressure inside the identification vessel should not exceed 50 psi gauge. For safety, the vessel

4.6 Apparatus Capabilities

104

pressure should not exceed the maximum pressure of 100 psi. The pressure transducers have
been calibrated from 0 to 50 psi, therefore readings are only accurate within these bounds. For
all tests a safety shut-off, or hydraulic interlock, was set to initiate once a pressure of 50 psi was
exceeded. It is recommended that this precaution be taken for all future tests conducted with the
apparatus.
Accuracy is also limited in the frequency domain. Inherent flexibility in the apparatus
induced a dominant mode of vibration at approximately 70 Hz, which altered the chamber pressure readings and skewed measurements. However, this limitation did not impede experimental
studies since all tests were conducted below 70 Hz.

The following chapter will present the

specific test configurations and system identification techniques used to extract hydraulic mount
parameters.

Chapter

Parameter Identification
The experimental apparatus presented in Chapter 4 is now used to identify the model parameters
established in Chapter 2. Throughout this chapter each main component within the hydraulic
mount system is studied in isolation.

For each case, the isolation method is presented and

system identification techniques are documented. Enough information is provided to enable the
extension of this work to encompass other mount designs. The design under investigation in this
chapter is the Cooper-Standard mount shown in Figure 5.1.

5.1

Upper Chamber Compliance

During the lumped parameter assignments in Section 2.1.1, the upper compliance was given four
parameters kr , br , Ap and C1 representing the linear stiffness, damping, effective pumping area
and volumetric compliance, respectively. To enhance these characteristics, the model developments in Section 2.3.1 assigned a resistance parameter to the bulge damping effect and made
all four stiffness and compliance parameters nonlinear. The following section presents the approach used to sort through each upper compliance parameter, beginning with the real stiffness
and damping characteristics.

105

5.1 Upper Chamber Compliance

106

Upper
Compliance

Decoupler

Inertia
Track
Lower
Compliance

Figure 5.1: Cross section of the Cooper-Standard mount under investigation.

5.1.1

Stiffness and Damping Parameters

The nonlinear stiffness kr and damping br parameters are defined to be a function of excitation
driving frequency wdr , and preload force Fp . From (2.22), these parameters are
amplitude X,
written as,

Fp
kr wdr , X,

Fp
br wdr , X,
Since these parameters are only dependent on the conditions applied to the whole mount system,
the internal fluid system dynamics have no influence on the stiffness and damping parameters.
The approach to investigating these parameters is to remove the fluid from the hydraulic mount
and conduct tests on the upper compliance, effectively in isolation. Test data for the stiffness
and damping parameters are collected using the same configuration and procedure used for a
hydraulic mount, presented in Section 1.2.

107

Compressive Force (N)

5.1 Upper Chamber Compliance

Mount Displacement (mm)


Measurements

Figure 5.2: Static force versus displacement.


The static force-displacement curve is measured and plotted in Figure 5.2. This curve
illustrates the hysteresis inherent in rubber, which gives most of the materials damping properties. The dynamic characteristics are measured using steady-state sinusoidal excitations with
the preload force held constant at 1700 N, reflecting in-vehicle conditions, and summarized in
Table 5.1.
Figure 5.3 shows the measured nonlinear stiffness and damping parameters found using
the test conditions in Table 5.1. Since these parameters have been established over the excitation
amplitude and frequency ranges of interest and are independent of internal fluid dynamics, they
were included in a look-up table for future system model simulations.
Table 5.1: Test conditions used for dynamic characterization, at 1700 N preload.
P-P Amplitude (mm)
1
2
0.1
0.2
0.3

Frequency Range (Hz)


2 to 40
2 to 40
10 to 250
10 to 250
10 to 250

Frequency Step (Hz)


2
2
10
10
10

kr (N/mm)

108

br (N s/mm)

br (N s/mm)

kr (N/mm)

5.1 Upper Chamber Compliance

Frequency (Hz)

Frequency (Hz)

(a)
1 mm P-P Amp

(b)
2 mm P-P Amp

0.1 mm P-P Amp

0.2 mm P-P Amp

0.3 mm P-P Amp

Figure 5.3: Stiffness and damping properties measured at various excitation conditions: (a)
low frequency high amplitude; (b) high frequency low amplitude.

5.1.2

Effective Pumping Area Parameter

The nonlinear model for the effective pumping area Ap (Fp ) is dependent on the preload force.
As the mount is displaced under the preload force, the rubber deforms so that the effective piston
area changes. To investigate this relationship, the experimental apparatus shown in Figure 5.4
was used to isolate the upper compliance. By completely filling the vessel with fluid below the
compliance, ensuring no air pockets are present, the changes in displacement at the mount top
pushed fluid into the lower vessel chamber. Using pressure sensor feedback, the vessel pressure
was maintained at 5 psi.

As the mount was displaced in increments of 1 mm, the pressure

control repositioned the actuator to accommodate the fluid displaced. Measurements of mount
displacement and actuator position were tabulated and used to determine the effective area as a
function of mount displacement, shown in Figure 5.5. To obtain the appropriate effective area
parameter using the preload force, the static stiffness curve in Figure 5.2 was first used to obtain
the preload displacement. For the mount under investigation, the effective pumping area was
found to be 3065 mm2 at 1700 N preload.

5.1 Upper Chamber Compliance

109

Mount
Displacement

Xact
Upper
Compliance

Ps1
Pressure
Feedback

Actuator
Position

Fluid
Filled
Chamber

Ap (mm2)

Figure 5.4: Test configuration for measuring the effective area and volumetric compliance
parameters.

Mount Displacement (mm)


Measurements

Figure 5.5: Plot of effective pumping area versus mount displacement.

5.1 Upper Chamber Compliance

5.1.3

110

Volumetric Compliance Parameters

Once the stiffness, damping and effective pumping area parameters are established, the volumetric compliance parameters can be investigated. Two approaches are used to extract the C1
and R1 parameters. Tests are first conducted on the chamber in isolation. Then a simplified
hydraulic mount is tested and used in conjunction with the system model to establish the same
parameters.
Using the test configuration illustrated in Figure 5.4, the upper compliance is displaced by
the amount corresponding to the 1700 N preload. A known volume of fluid is then forced into
the upper chamber and pressure readings are collected. The static curve of volume displacement
versus pressure is shown in Figure 5.6. These measurements illustrate the bulge damping effect
from the rubber, which is modeled by introducing the resistance parameter R1 . Previous experimental studies documented by Kim and Singh [12] limit the investigation of compliance to static
analysis and neglect the effects of hysteresis. However, the present work includes experimental
studies on the dynamic volumetric compliance and bulge damping.
Dynamic characteristics of the volumetric compliance and bulge damping parameters,
C1 (wdr , VT , Fp ) and R1 (wdr , VT , Fp ) are investigated by establishing a local linear response for
various driving frequencies wdr and volume amplitudes VT . This is achieved by applying sinusoidal actuator excitations and measuring lower chamber pressure response.
The excitations are applied under a controlled mean pressure of 15 psi, and steady-state
time domain data is acquired for actuator position Xact and pressure Ps1 . The actual volume of
fluid passing into the upper chamber Vactual is then computed using equations (4.1) to (4.9). The
parameters C1 and R1 are identified through the compliance model developed in Section 2.3.1,
equation (2.24),

C1 P1 = QT + C1 R1 Q T

(5.1)

111

Volume (mm3)

5.1 Upper Chamber Compliance

Pressure (MPa)
Measurements

Figure 5.6: Static curve of volume versus pressure in the upper chamber.
For the isolated experimental tests the variables QT and P1 are replaced by the measured
system input and output variables Qactual and Ps1 , respectively, where Qactual is the derivative of
Vactual . Integrating (5.1) with respect to time and assuming all initial conditions are zero, gives
Ps1 =

Vactual
+ R1 V actual
C1

(5.2)

To cast equation (5.2) into the frequency domain, the steady-state pressure and volume variables
take the form
Ps1 = Ps1 ejwdr t

(5.3)

Vactual = Vactual ejwdr t


where Ps1 and Vactual represent the steady-state amplitude and phase of pressure and volume,
respectively. The final frequency domain equation becomes,
Ps1
1
=
+ R1 wdr j

C1
Vactual

(5.4)

A Fourier transformation of steady-state Vactual and Ps1 measurements is used to establish


the real and imaginary components of equation (5.4). The parameters C1 and R1 are identified
using the measured frequency domain data and are illustrated in Figure 5.7.
A few points should be noted regarding this test configuration. First, the frequency range is
limited to under 70 Hz due to the limited capabilities of the test apparatus. Therefore, volumetric

5.2 Lower Chamber Compliance

112

compliance parameters are not identified at high frequencies. In addition, the test configuration
fixes the upper compliance so that the normal excitation point is stationary. Since this point is
in motion during excitations, the test setup does not exactly represent the physical system.
To address the above mentioned issues, the C1 and R1 parameters are also identified using
a special case hydraulic mount.

A mount designed with a long narrow inertia track and no

decoupler is used to extract the compliance parameters. The inertia track geometry generates
high inertia and damping properties which result in a liquid column resonance below 20 Hz.
Figure 5.8 illustrates that the inertia track has negligible influence on the system above 30 Hz.
Since the stiffness, damping and effective pumping area parameters are already identified, the
model response in the high frequency range is tuned by adjusting the C1 and R1 parameters
to match the dynamic stiffness characteristics of the special mount.

With this approach the

volumetric parameters are identified from 30 Hz to 250 Hz and illustrated in Figure 5.9.
The upper compliance parameters in (2.25) are dependent on the fluid volume amplitude
being pushed into the upper chamber. Since this volumetric expansion is a function of internal
fluid dynamics, the C1 and R1 parameters cannot be directly calculated at each excitation condition. To establish parameter values during the model simulations, an iterative procedure is
applied to update compliance parameters based on the steady-state volume response. Using an
initial guess of C1 and R1 , the steady-state simulation response is used to calculate volume flow
amplitude VT into the upper chamber. The compliance parameters are updated using a surface
fit to the data in Figure 5.9, then the simulation is reiterated. Complete details of the simulation
procedure are presented in Chapter 6.

113

C1 (mm /N)

R1 (Ns/mm5)

5.2 Lower Chamber Compliance

Frequ

ency (H

m
Volu

z)

e Am

ude
p lit

(m m

Frequency
(H

(a)

me
Volu

z)

de
litu
Amp

(m m

(b)

Phase (Degrees)

Phase (Degrees)

Dynamic
Stiffness (N/mm)

Dynamic
Stiffness (N/mm)

Figure 5.7: Upper compliance parameters identified at 15 psi mean pressure using the test
apparatus: (a) volumetric compliance; and (b) bulge damping.

Frequency (Hz)

Frequency (Hz)

(a)
1 mm P-P Amp

(b)
2 mm P-P Amp

0.1 mm P-P Amp

0.2 mm P-P Amp

0.3 mm P-P Amp

Figure 5.8: Dynamic stiffness and phase response of the special mount, measured using the
test conditions in Table 5.1: (a) low frequency; and (b) high frequency response.

114

C1 (mm /N)

R1 (Ns/mm )

5.2 Lower Chamber Compliance

Frequ
ency (
Hz)

um
Vol

m
e (m
itud
l
p
m
eA

(a)

Frequency
(Hz)

me
Volu

(
ude
plit
Am

)
mm

(b)

Figure 5.9: Upper compliance parameters identified using the system model and test conditions
in Table 5.1: (a) volumetric compliance; and (b) bulge damping.

5.2

Lower Chamber Compliance

The mount under investigation (see Figure 5.1) is considered to be a single-load-bearing-chamber


(SLBC) mount, since the lower chamber serves mainly as a fluid transfer area [6]. Essentially,
the lower chamber compliance acts as a diaphragm, contributing little to the volumetric stiffness
of the system. However, the assignment of a compliance parameter, C2 (Fp ) (2.26), to the lower
chamber is still carried out. This parameter is only dependent on the preload force, or the mean
volume of fluid in the lower chamber.
To isolate the C2 parameter, the test apparatus was used with the lower chamber fastened
across the holding plate, as illustrated in Figure 5.10. The static volume versus pressure curve
was measured and is plotted in Figure 5.11.
During model simulation the C2 parameter is calculated by first estimating the volume of
fluid displaced into the lower chamber under the preload force. This is done using the effective
pumping area curve in Figure 5.5 and the static stiffness curve in Figure 5.2.

The slope of

Figure 5.11 at this volume level is used to approximate the volumetric compliance of the lower

5.2 Lower Chamber Compliance

Lower
Compliance

115

Xact
Actuator
Position

Ps1
Pressure

Fluid
Filled
Chamber

Volume (mm3)

Figure 5.10: Test configuration for determining the volumetric compliance of the lower chamber.

Pressure (MPa)
Measurements

Figure 5.11: Static measurements of lower chamber volumetric expansion versus pressure.

5.3 Inertia Track Parameters

116

chamber. It should be noted that some hydraulic mounts are filled after the preload force is
applied, thus the volume displaced is different than that mentioned here.
For the mount under investigation, the volume displaced was found to be 32944 mm3 at
the 1700 N preload, and the resulting lower chamber volumetric compliance was calculated to
be 2.6 106 mm5 /N.

5.3

Inertia Track Parameters

Inertia track parameters are identified experimentally using the test apparatus configuration shown
in Figure 5.12. The decoupler and inertia track assembly is bolted to the holding plate within the
identification vessel. By blocking the decoupler, the fluid flow across the holding plate V actual
becomes the inertia track flow and transducer measurements Ps1 and Ps2 form the pressure differential in the momentum equation. To ensure that pressure levels within the vessel never exceed
the 0 to 50 psi range, the vessel is sealed with an air pocket and pressurized to a mean operating level of 20 psi. Using pressure feedback on Ps2 , the upper vessel chamber is controlled to
operate at the desired mean level during all test conditions.
The nonlinear inertia track model developed in Section 2.3 equation (2.29) is
P1 P2 = Ii Q i + (Ri1 + Ri2 |Qi |) Qi
where the parameters Ii , Ri1 and Ri2 are first investigated using frequency sweep data and a
simplified linear momentum equation. Sinusoidal excitations are applied to the actuator and
steady-state time domain data is acquired of the actuator position Xact and both chamber pressures Ps1 and Ps2 . Fluid volume passing through the inertia track Vactual is then calculated and
used with the pressure differential Ps1s2 = Ps1 Ps2 in the isolated system equation. The
isolated response follows a linearized form of equation (2.29),
Ps1s2 = Ii Vactual + Ri V actual

(5.5)

5.3 Inertia Track Parameters

117

Cap to
seal vessel
X

Inertia
Track

Fluid
Fill
Level

Actuator
Position

Ps2
Ps1
Pressure
Sensors

Figure 5.12: Test configuration for isolating the inertia track paramters.
This equation is transformed to the frequency domain using the steady-state solutions,
Ps1s2 = Ps1,s2 ejwdr t

(5.6)

Vactual = Vactual ejwdr t


where Ps1,s2 and Vactual represent the complex pressure differential amplitude and volume amplitude, respectively. The linear parameters are then identified using
Ps1s2
2
= Ii wdr
+ Ri wdr j
Vactual

(5.7)

Measured frequency response data is calculated by applying the Fourier transform on steady-state
time domain data of Vactual and Ps1s2 . The Ii and Ri parameters are plotted in Figure 5.13
for several excitation conditions. The trends confirm that the fluid column inertia remains constant regardless of flow through the inertia track and suggest that the resistance parameters are
dependent on the flow, as indicated by the nonlinear momentum equation (2.29).

5.3 Inertia Track Parameters


x 10

-6

x 10

118

-4

Resistance (kg/s-mm4)

Inertia (kg/mm4)

2 .5

1 .5

1 .5

0 .5
0 .5

0.5

1.5

2.5

3.5

0.5

1.5

2.5

Flow (mm /s)

x 10

3.5

4
5

Flow (mm /s)

(a)

x 10

(b)

Figure 5.13: Parameter trends with respect to inertia track flow, for: (a) inertia; and (b) resistance.
The resistance parameters in equation (2.29) are approximated using the trendline shown
in Figure 5.13b where Ri1 is the y intercept and Ri2 is the slope. For the inertia track under
investigation, the parameters are determined to be
Ii = 3.29 106 kg/mm4
Ri1 = 4.9 105 kg/s-mm4
Ri2 = 4.6 1010 kg/mm7
Although this approach is sufficient for determining the nonlinear inertia track parameters,
a more elegant method can be applied which significantly reduces testing time. To introduce this
method, equation (2.29) is rewritten in terms of the measured experimental variables.
Ps1s2 = Ii Q actual + (Ri1 + Ri2 |Qactual |) Qactual

(5.8)

where Qactual represents the time derivative of Vactual . For each time sample of pressure differential and flow, equation (5.8) defines the relationship between variables within the isolated
system. In theory, this suggests that three samples of measured data is enough to set up three

5.3 Inertia Track Parameters

119

equations and solve each of the desired parameters. If a sequence of time domain data is acquired, the least squares method will determine the best estimate of system parameters. This
approach first requires the sequence of measurements be placed in matrix form, as follows

1
1
Qactual 1 Qactual 1 |Qactual | Qactual

Ps1s2
I

2
2
2
i
2 Ps1s2
Qactual Qactual
|Qactual | Qactual

=
(5.9)

Ri1
..
..
..
..

R
.
.
.
.
i2
n
n
Ps1s2
Qactual n Qactual n |Qactual | Qactual
By assigning each matrix to the notation,

1
Qactual
Ps1s2
2

2
Ps1s2
Q actual
; U =
Y =

.
..

..

.
n
n

Ps1s2
Qactual

Qactual
2
Qactual
..
.

Qactual

|Qactual | Qactual
2
|Qactual | Qactual
..
.
|Qactual | Qactual

the least-squares parameter estimation is then applied using

1 T
= U T U
U Y

Ii
= Ri1
Ri2

(5.10)

where is the least squares estimate of the parameters in , (see Wilson [34] or Juang [9])
To apply this technique, a random perturbation is applied to the actual volume input as
illustrated in Figure 5.14a. The frequency spectrum of the excitation is measured and plotted in
Figure 5.14b, which indicates whether the perturbation is within the frequency range of interest.
The random excitation is applied for 10 seconds and sampled every 0.002 seconds. Data samples
are assembled into the form presented in (5.9) to obtain the inertia track parameters and the
estimated output illustrated in Figure 5.15. Parameter estimation using this technique identifies
Ii = 3.38 106 kg/mm4
Ri1 = 6.52 105 kg/s-mm4
Ri2 = 4.49 1013 kg/mm7
These results indicate that both the sine sweep and random least squares method can be applied
successfully to extract the inertia track parameters. Since the random perturbation method is

5.3 Inertia Track Parameters

120

11

Flow (mm3/s)

Flow
6 2
Spectral Magnitude (mm /s )

x 10
3
2.5
2
1.5
1
0.5
0

10

15

20

25

30

Frequency (Hz)

Time (s)
(a)

(b)

Figure 5.14: Isolated inertia track excitation plots including: (a) time domain segment of random flow input; and (b) power spectral density of complete flow perturbation.

Pressure Differential (MPa)

0.1

0.05

-0.05

-0.1
2.05

2.1

2.15

2.2

2.25

2.3

2.35

2.4

2.45

2.5

Time (s)
Measured

Least squares estimate

Figure 5.15: Time domain segment of measured pressure differential versus model output
using estimated parameters.

5.4 Decoupler Parameters

121

experimentally faster and computational less intensive than the sine sweep method, it is recommended for inertia track parameter identification.

5.4

Decoupler Parameters

The experimental test configuration used to identify the decoupler parameters is shown in Figure 5.16. To ensure that all flow forced across the holding plate passes through the decoupler,
the inertia track is blocked. The actual volume Vactual is used with the output pressure differential Ps1s2 to identify the system parameters through the momentum equation. Pressure levels
within the vessel are maintained at a mean operating level of 20 psi to ensure the transducer
limits are not exceeded.
The nonlinear decoupler model established in Section 2.3 equations (2.36) and (2.59) include a nonlinear momentum equation in parallel with a variable resistance orifice,

P1 P2 = Id Q d + Rd1 + Rd2 |Qd | + 1 eRd3 Xd arctan(Qd 2 ) Qd


P1 P2 = RL2_f nc |QL | QL

respectively. Five parameters in equations (2.36) and (2.59) control the complete behavior of
the nonlinear decoupler model. The inertia Id and resistances Rd1 and Rd2 govern the decoupler
properties when operating in the uncoupled region, while Rd3 controls the decoupler switching. Also, the nonlinear leak flow resistance RL2_f nc , exists as RL2 when the decoupler is closed
and is controlled with the parameter Vd_ max .

Even with the system experimentally isolated,

equations (2.36) and (2.59) represent a complex five parameter identification task with one state
variable Vd . As a consequence, the following approach utilizes a two step identification procedure.
The first experimental test is a frequency sweep, similar to the inertia track test. If the
excitation volume amplitude is such that the decoupler is kept within the uncoupled region, the

5.4 Decoupler Parameters

122

Cap to
seal vessel
Xact
Fluid
Fill
Level

Decoupler

Actuator
Position

Ps2
Ps1
Pressure
Sensors

Figure 5.16: Test configuration for isolating the decoupler parameters.


system response will follow equation (2.31)
P1 P2 = Id Q d + (Rd1 + Rd2 |Qd |) Qd
which is a reduced form of equation (2.36). Simplifying equation (2.31) into a linear form with
the experimentally measured variables gives,
Ps1s2 = Id Vactual + Rd V actual

(5.11)

This linear momentum equation is then converted to the frequency domain using the steady-state
expressions in (5.6).
Ps1s2
2
= Id wdr
+ Rd wdr j

Vactual

(5.12)

Equation (5.12) is used to identify the inertia and resistance parameter trends, as conducted
for the inertia track. However, the important distinction to this frequency domain test is the
dependance on volume amplitude. For all excitation conditions, equation (5.12) only applies if
the response is linear, or the decoupler does not contact its cage. For this reason, each steady-

123

Pressure (MPa)

Pressure (MPa)

Volume (mm )

Volume (mm5)

5.4 Decoupler Parameters

Time (s)

Time (s)

(a)

(b)

Figure 5.17: Steady-state time domain data of actual decoupler volume and pressure differential at 12 Hz, showing: (a) nonlinear cage contact; and (b) linear uncoupled response.
state response is first checked to ensure the pressure differential does not indicate decoupler cage
contact. Figure 5.17 illustrates the desired characteristics of steady-state response data.
The Fourier transform is applied to the steady-state data to establish the frequency response
characteristics and identify Id and Rd using equation (5.12). Identified parameters are plotted
over the decoupler flow amplitude in Figure 5.18.
The isolated response demonstrates that the decoupler inertia remains constant across various flow conditions, while the resistance increases with the decoupler flow; validating the nonlinear uncoupled equation (2.31). A reasonable approximation for parameters Rd1 and Rd2 is
established using the intersection and slope of the linear trendline. For the mount under investigation, the parameters are found to be

Id = 8.28 108 kg/mm4


Rd1 = 4.58 106 kg/s-mm4
Rd2 = 4.25 1011 kg/mm7
The amplitude-sensitive behavior of the decoupler gives unreliable results using the least
squares method. Therefore, the sine sweep technique provides the best method to identifying
parameters pertaining to the uncoupled decoupler.

5.4 Decoupler Parameters


x 10

-8

x 10

124

-6

Resistance (kg/s-mm )

Inertia (kg/mm4)

7
6

4
3

Flow (mm3/s)

Flow (mm3/s)

x 10

(a)

8
4

x 10

(b)

Figure 5.18: Parameter trends with respect to decoupler flow, for: (a) fluid column inertia; and
(b) resistance.
The two remaining parameters Rd3 and RL2 combine with the decoupler state Vd , or unknown plate position Xd , to form a state and parameter identification problem. Since the switching characteristics under investigation are independent of excitation frequency, the second isolated test is conducted by applying low frequency random perturbations. Under these excitations the decoupler inertia has negligible impact on the isolated system response, therefore it is
removed from equations (2.36) and (2.59). Also, the nonlinear leak flow resistance function
RL2_f nc is reduced to the RL2 parameter, since the flow is controlled when isolated in the test
apparatus. The isolated switching model becomes,
Ps1,s2 =

Rd1 + Rd2 |Qd | + 1 eRd3 Xd arctan(Qd 2 ) Qd

Ps1,s2 = RL2 |QL | QL

(5.13)
(5.14)

where 1 and 2 are maintained at constant values of 1 1016 and 1 105 , respectively, as
established in Section 2.3. Also, recall from equation (2.33) that
Xd =

Vd
Ad

5.4 Decoupler Parameters

125

where the decoupler area Ad is measured from geometry of the mount under investigation.
In equations (5.13) and (5.14) the actual flow controlled across the decoupler Qactual now
represents the total volume accommodated via the decoupler and leak flow,
Qactual = QL + Qd

(5.15)

Also, the absolute value signs in equations (5.13) and (5.14) are resolved into algebraic equations
by noting the signs of Qd and QL are equivalent to that of the input Qactual . Using the arctangent
function presented in Section 2.3, the absolute values of decoupler and leak flow are transformed
according to:
2
Qd arctan(Qactual )

2
(Qactual Qd ) arctan(Qactual )
|Qactual Qd | =

|Qd | =

(5.16)
(5.17)

Using (5.15) to (5.17), the equations (5.13) and (5.14) are transformed to
2
Ps1s2 = Rd1 Qd + Rd2 Q2d arctan(Qactual ) + 1 eRd3 Xd arctan(Qactual 2 ) Qd

2
Ps1s2 = RL2 (Qactual Qd )2 arctan(Qactual )

(5.18)
(5.19)

Equations (5.18) and (5.19) are equated to form a function for decoupler flow Qd , by solving the
quadratic equation

2
0 =
(RL2 Rd2 ) arctan(Qactual ) Q2d
(5.20)

4
Rd3 Xd arctan(Qactual 2 )
Rd1 Qd + 1 e
+ RL2 Qactual arctan(Qactual ) Qd

2
+RL2 Q2actual arctan(Qactual )

where Xd is proportional to the state variable Vd , through (2.33).

The equation for Qd , is

simplified into a function of the state Vd , parameters Rd3 and RL2 , and input Qactual
Qd = f (Vd , Rd3 , RL2 , Qactual )

(5.21)

5.4 Decoupler Parameters

126

Using the decoupler flow (5.21) in equation (5.19) gives the system output equation, also written
in reduced form as
Ps1s2 = g(Vd , Rd3 , RL2 , Qactual )

(5.22)

Equations (5.21) and (5.22) form the state and parameter identification problem, where
the decoupler state has to be predicted and each parameter estimated, for each position in time.
Convergence issues arise when attempting to predict Vd and identify Rd3 and RL2 all in one
algorithm. Therefore, RL2 is calculated prior to predicting the state variable.
The leak flow resistance is first calculated using the measured pressure differential output,
as illustrated in Figure 5.19. The peak pressure differential measurements (circled locations on
Figure 5.19) represent the high resistance leak flow around the closed decoupler, occurring when
QL = Qactual . The RL2 parameter is then calculated using
RL2 =

Ps1s2
Q2
actual

(5.23)

where Ps1s2 and Qactual represent the measured data at the pressure differential peaks.
Once RL2 is estimated, the functions in (5.21) and (5.22) are reduced to just one unknown
parameter Rd3 , and the state variable Vd .
Qd = f (Vd , Rd3 , Qactual )

(5.24)

Ps1s2 = g(Vd , Rd3 , Qactual )

(5.25)

Equations (5.24) and (5.25) are discretized, using


Qd,k =

Vd,k+1 Vd,k
T

(5.26)

where k represents the k th sample point of the data series, sampled at time intervals T . Using
(5.26) to transform equations (5.24) and (5.25) into discrete form gives,
Vd,k+1 = f (Vd,k , Rd3,k , Qactual,k , k)

(5.27)

Ps1s2,k = g(Vd,k , Rd3,k , Qactual,k , k)

(5.28)

127

Pressure (MPa)

Flow (mm5/s)

5.4 Decoupler Parameters

Time (s)

Figure 5.19: Time segment of the decoupler input flow and pressure differential, indicating
locations where leak resistance is identified.
Identification of both states and parameters presents a unique problem. First, the state
estimation problem formulation assumes knowledge of all system parameters. The system identification problem to determine parameters then assumes the availability of the variables or states
within the system. To work around this problem the parameter Rd3 is written as a state variable and appended to the original state (see Wilson [34]). Since Rd3 is constant, the new state
equation becomes
Rd3,k+1 = Rd3,k

(5.29)

this state is then added to Vd,k+1 to form the new state equation vector
k+1 =
X

Vd,k+1
Rd3,k+1

f(Vd,k , Rd3,k , Qactual,k , k)


Rd3,k

(5.30)

and the state vector


k =
X

Vd,k
Rd3,k

(5.31)

5.5 Bell Parameters

128

Therefore, the new isolated decoupler system becomes


k+1 = f (X
k , Qactual,k , k)
X

(5.32)

k , Qactual,k , k)
Ps1s2,k = g(X

(5.33)

The new state vector (5.31) transforms the system into a nonlinear state estimation problem, which is then solved using the Extended Kalman Filter
k+1 = f (X
k , Uk , k) + wk
X
k , Uk , k) + qk
Zk = g(X

(5.34)
(5.35)

where Uk and Zk represents the system input and output, respectively. The wk denotes disturbance noise on the states, while qk represents measurement noise on the output. The complete
Extended Kalman Filter algorithm used to identify state variables from equations (5.34) and
(5.35), is presented in Appendix D.
Random perturbations are applied to the isolated system for approximately 20 seconds and
sampled at time intervals T = 0.004 s. Input and output measurements are used to solve for
RL2 , then analyzed using the filter algorithm. The results in Figure 5.20 show how the procedure
converges on parameter estimates, while Figure 5.21 provides a closer look at the predicted state
motion and output accuracy. For the decoupler under investigation this identification procedure
obtained
RL2 = 9.21 109 kg/mm7
Rd3 = 14.5 mm1
The Vd_ max parameter, used as the switching parameter in the nonlinear transmitted force
(2.52) and leak flow (2.58) equations, is determined using the converged decoupler state estimate
Vd . As Figure 5.21b indicates, Vd_ max has an estimated value
Vd_ max = 750 mm2

5.5 Bell Parameters

0.06
5

Volume (mm )

2000

0.04

Pressure (MPa)

129

0.02
0

1000
0
-1000
-2000

-0.02

10

15

20

15

20

(b)

19.2
-0.04

Rd3

16.0
-0.06

12.7

-0.08
-0.1

9.5
0

10

15

3.2

20

10

Time (s)

Time (s)

(a)

(c)

Figure 5.20: Decoupler state and parameter identification: (a) pressure differential across decoupler; (b) decoupler volume state variable; and (c) decoupler switching parameter.

1000

Volume (mm )

0.05

Pressure (MPa)

0.04
0.03
0.02

500
0
-500
-1000

0.01

15

16

17

18

19

18

19

(b)

0
-0.0

16.0

Rd3

-0.02

12.7

-0.03
9.5

-0.04
15

16

17

18

Measured

19

15

16

17

Time (s)

Time (s)
Estimate

(c)

(a)

Figure 5.21: Close up of the decoupler state and parameter identification: (a) pressure differential across decoupler; (b) decoupler volume state variable; and (c) decoupler switching parameter.

5.5 Bell Parameters

5.5

130

Bell Parameters

The bell system is also experimentally studied using the same Cooper-Standard mount. A bell is
included in the upper chamber such that when the 1700 N preload is applied, the bell operates 5
to 10 mm from the decoupler plate, as illustrated in Figure 5.22. In the current Cooper-Standard
mount, the operating position of the bell is such that the bell chamber compliance is equivalent
to the previously measured upper chamber. To obtain full knowledge of all parameters, this
configuration only requires the identification of the bell fluid column inertia, resistance, and new
upper chamber compliance.

5.5.1

Fluid Column Inertia and Resistance

The mathematical model developed to represent the bell system in Section 2.4 assigns inertia
and resistance parameters to the annular column of fluid between the bell plate and mount wall.
These parameters are now experimentally measured using the test apparatus configuration illustrated in Figure 5.23. A section of the mount wall is clamped into the holding plate within the
apparatus. The bell plate is then fixed into position between the wall segment with a 12 -in wide
bracket. To control fluid flow through the bell, the bracket is designed so that it does not obstruct
flow around the bell plate. The fluid volume Vactual pushed across the holding plate becomes the
bell flow, while the pressures Ps1 and Ps2 are measured. As with the inertia track and decoupler
tests, the vessel is sealed with an air pocket and pressurized to a mean operating level of 20 psi.
The nonlinear bell momentum from equation (2.81) is

P1 Pb = Ib Qb + Rb1 (Qb + Ab X) + Rb2 Qb + Ab X (Qb + Ab X)

where P1 Pb is the pressure differential across the bell plate, equivalent to the measured differential in the isolated case. The variable X represents the bell plate motion in the system model,
and is zero for the isolated test configuration. Therefore, equation (2.81) can be simplified to a

5.5 Bell Parameters

131

Upper
Compliance
Bell Plate
Bell Position
With Preload

Decoupler

Inertia
Track
Lower
Compliance

Figure 5.22: Cross section of Cooper-Standard mount that includes a bell plate.

Cap to
seal vessel

wide
bracket
Segment of
mount wall

Fluid
Fill
Level

Xact
Actuator
Position

Bell Plate

Ps2
Ps1
Pressure
Sensors

Figure 5.23: Test configuration for isolating the bell parameters.

5.5 Bell Parameters

132

nonlinear momentum relation using measured variables:


Ps1 Ps2 = Ib Q actual + (Rb1 + Rb2 |Qactual |) Qactual

(5.36)

Equation (5.36) is now simplified to the linear momentum equation,

Ps1s2 = Ib Vactual + Rb V actual


similar to the isolated inertia track and decoupler tests. The linear moment is converted to the
frequency domain, using the steady-state expressions in (5.6),
Ps1s2
2
= Ib wdr
+ Rb wdr j

Vactual

(5.37)

This expression is used to identify the inertia and resistance parameters from the Fourier
transforms of the steady-state measurements. The identified parameters are plotted over the bell
flow amplitude in Figure 5.24. These experimental results also demonstrate a constant fluid
column inertia over the complete flow range, while increases in resistance are evident with increased bell flow. The trends validate the nonlinear momentum equation (5.36), with reasonable
approximations for parameters Rb1 and Rb2 established using the intersection and slope of the
linear trend line. For the mount under investigation, the parameters were found to be
Ib = 2.27 108 kg/mm4
Rb1 = 2.33 108 kg/s/mm4
Rb2 = 2.97 1012 kg/mm7

5.5.2

Chamber Compliance

The position of the bell was selected such that the bell compliance, or chamber above the bell, is
the upper compliance already investigated in Section 5.1, leaving the middle chamber unknown.

5.6 Summary
x 10

133

-8

x 10

-6

1 .2

Resistance (kg/s-mm4)

Inertia (kg/mm4)

1 .5

0 .8

0 .6

0 .4

0 .5
0 .2

1.5

2.5

3.5

Flow (mm3/s)

1.5
5

x 10

(a)

2.5

Flow (mm3/s)

3.5
5

x 10

(b)

Figure 5.24: Parameter trends with respect to bell flow, for: (a) inertia; and (b) resistance.
As illustrated in Figure 5.22, the chamber between the bell plate and inertia track - decoupler
housing assembly, referred to now as the upper chamber, will contain negligible bulge damping
and very little volumetric compliance C1 .
To approximate C1 , it is assumed that the steel housing deflects under pressure and gives
the chamber some volumetric compliance. The inertia track - decoupler housing is modeled
as a 79-mm diameter and 1.8-mm thick steel disk, and finite element methods are applied to
approximate the plate compliance.

A quarter plate element model, shown in Figure 5.25a,

illustrates the constraints and distributed pressure load. The linear deflection in Figure 5.25b
is used to calculate the volume displaced under the pressure applied. Using this approach, the
volumetric compliance of the upper chamber is found to be
C1 = 680 mm5 /N
It should be noted that although this volumetric compliance value is in the order of one tenth
that of the bell, it cannot be assumed rigid. The stiffness between the bell and decoupler flow is
necessary to establish the coupled multi DOF system.

5.6 Summary
Pressure Load

134

3.
3.

3.
3.
3.
3.
3.
3.
3.

3.
F

3.

F
3.
F

3.

24

3.

24

3.

3.

24

3.

24

3.

24

3.

24
24

F
15

24
24
24

3.

24
15

15

15

24
15

15

15

24
15

15

15

15

24
15

15

15

15

24
1245

Constraints

(a)

(b)

Figure 5.25: Illustration of the quarter plate model used to approximate the chamber compliance, showing: (a) boundary constraints and 3 MPa evenly distributed pressure load; (b) plate
deflection in linear elastic region.

5.6

Summary

This chapter has provided a systematic look at measuring each of the mount components in
isolation. Each of the upper chamber compliance parameters were identified through one of
three tests. The linear stiffness and damping were isolated by testing a dry mount and the
effective pumping area was isolated in the testing fixture and plotted as a function of mount
displacement.

Finally, the best method to extract the upper chamber volumetric compliance

parameters was found using a mount with a long narrow inertia track and no decoupler. The
lower chamber volumetric compliance characteristics were only considered to be a function of
preload conditions and were measured using the test apparatus.
The later portion of this chapter focused on the inertia track, decoupler and bell system.
Isolating these components in the experimental fixture enabled direct measurement of flow and
pressure differential, which were then used in the system identification procedures to extract the
parameters. For the inertia track, the inertia and resistance parameters were best extracted using
the least-squares method with random flow perturbations. The decoupler involved a two phase
identification procedure. First, a frequency sweep test was conducted to establish trends in the
uncoupled inertia and resistance. Large amplitude random excitations were then used in the

5.6 Summary

135

Extended Kalman Filter to extract the switching parameter and leak flow resistance. Finally,
the liquid column inertia and resistance associated with the bell were extracted using frequency
sweep measurements. In each of these cases the parameters identified pertained to the CooperStandard mount illustrated in Figure 5.22 and are summarized in Table 5.2.
Table 5.2: Inertia track, decoupler and bell parameters identified for the Cooper-Standard
mount.
Ii
Ri1
Ri2
Id
Rd1
Rd2

=
=
=
=
=
=

3.38 106 kg/mm4


6.52 105 kg/s-mm4
4.49 1010 kg/mm7
8.28 108 kg/mm4
4.58 106 kg/s-mm4
4.25 1011 kg/mm7

RL2
Rd3
Ib
Rb1
Rb2

=
=
=
=
=

9.21 109 kg/mm7


14.5 mm1
2.27 108 kg/mm4
2.33 108 kg/s-mm4
2.97 1012 kg/mm7

Although the modeling approach throughout this thesis has been to capture the mount
characteristics in as much detail as possible, it may be feasible to reduce model complexity in
some areas.

Specifically, while the inertia track measurements indicate a strong correlation

with the nonlinear resistance model, the decoupler results do not present such a strong case.
The relative magnitude and impact of the free decoupler resistance parameters suggest that the
resistance could perhaps be simplified to a linear form without significant degradation in the
overall system response. However, this is not investigated further as this thesis remains focused
on establishing a complete hydraulic mount model.

Chapter

Simulation and Validation of System Models


In this chapter, the nonlinear models developed in Chapter 2 are simulated using the parameters
identified in Chapter 5. Simulation procedures are explained for the inertia track and decoupler
model, as well as the additional bell system.

The model response is then compared with

measured data using dynamic stiffness and phase responses. The models are compared over
the complete frequency and amplitude ranges of interest to validate the modeling and parameter
identification techniques.

6.1

Simulation Techniques

The simulation techniques used to obtain frequency domain data from the nonlinear inertia track
and decoupler model, are presented first. Dynamic system equations, summarized in Section 2.7,
are rewritten in state-space equation form. The continuity equations are
1

Ap X Qi Qd QL
+ R1 Ap X Qi Qd
C1
1
= (Qi + Qd + QL )
C2

P1 =

(6.1)

P2

(6.2)

and the momentum equations become


1
Q i = (P1 P2 (Ri1 + Ri2 |Qi |) Qi )
Ii

1
Q d = P1 P2 Rd1 + Rd2 |Qd | + 1 eRd3 Xd arctan(Qd 2 ) Qd
Id
136

(6.3)
(6.4)

6.1 Simulation Techniques

137

An additional state equation is required for the decoupler volume,


V d = Qd

(6.5)

which is then related to the decoupler displacement via


Xd =

Vd
Ap

The leak flow across the closed decoupler, is written as


s
2 |(P1 P2 )|
QL =
arctan ((P1 P2 ) 8 )

RL2_f nc

(6.6)

(6.7)

with the nonlinear resistance equation (2.58)

2
2
Vd arctan( 5 Qd ) Vd_ max 6
RL2_f nc = RL2 1 + 7 7 arctan

Equations (6.1) to (6.7) represent a system of first order differential equations which can
be solved using numerical techniques.

However, input conditions must be considered.

In

the nonlinear model development and parameter identification methods, all variable integration
assumed initial conditions of zero.

Therefore, the model simulations require that all initial

jwdr t
conditions be set equal to zero. Since the direct application of input excitation X = Xe
a simulated actuator is required to generate
will give non-zero initial values of both X and X,
the desired input with the appropriate initial conditions.
The single DOF actuator modeled in Figure 6.1 consists of a mass, held in place with a
spring and dash-pot. As illustrated, a force input F (t) on the mass will produce the desired
displacement output X (t) through the equation of motion
+ cX + kX = F
mX

(6.8)

where m, c and k represent the mass, damping and stiffness parameters, respectively. To ensure
the actuator output X(t) is not influenced by the mount being excited, the value of the stiffness
parameter k is several orders of magnitude larger than a typical mount stiffness. This effectively
makes the mount stiffness irrelevant, hence it is not included in equation (6.8). The mass is

6.1 Simulation Techniques

138

F(t)

k
m

X(t)

Figure 6.1: Actuator model used to provide excitations with zero initial conditions.
calculated using
m=

k
wn2

(6.9)

where wn is the actuator natural frequency, selected to be 5000 Hz, well above the frequencies
being excited. The damping parameter is selected to achieve critical damping,
c = 2wn m

(6.10)

although it is perhaps not necessary, since the natural frequency is so far beyond the frequency
range of interest. The forced input amplitude F is calculated using the steady-state frequency
response

mw2 + jcwdr + k ejwdr t = F ejwdr t


X
dr

(6.11)

the time domain input force is,


To achieve the desired amplitude of excitation X,

2
mwdr
F (t) = X
+ jcwdr + k ejwdr t

(6.12)

Writing equation (6.8) in state space form, gives

X = X

= F kX cX 1
X
m

(6.13)
(6.14)

6.1 Simulation Techniques

139

which is added to the existing differential equations to form a series of seven state equations.
Since the equations (6.1) to (6.5) and (6.13) to (6.14) are in the form y = f (t, y), a numerical
solution can be obtained over a time interval [to , tf ] given initial values yo (to ). For each frequency and amplitude of excitation, the state equations are solved for 10 cycles and sampled at
200 points per cycle. Since the decoupler model introduces small time constants, as illustrated
in Figure 3.10, the equations are considered to be Stiff.

Using MATLAB [19] the set of

equations are solved by applying a new numerical differentiation technique for stiff differential
equation sets based on Bogacki-Shampine [14]. The numerical differentiation returns the time
domain solution of each state variable in the system. Once the differential equations are solved
, the transmitted force equation is calculated using the time domain state variable solutions, and
equation (2.53)
FT = kr X + br X + (Ap Ad_f nc )(P1 P2 )
+Ap P2 + (Ad Ad_holes ) (Rd1 + Rd2 |Qd |) Qd
where equation (2.52) gives
Ad_f nc

1
= Ad

arctan
2


2
Vd arctan ( 3 P ) Vd_ max 4

Before establishing the dynamic stiffness characteristics, the compliance C1 (wdr , VT , Fp )


and bulge damping R1 (wdr , VT , Fp ) parameters must be correlated with the volume amplitude of
fluid entering the chamber VT . Volume flow into the upper chamber is calculated by integrating
internal flows
VT =

Ap X Qd Qi QL dt

(6.15)

The amplitude of fluid entering the upper chamber VT is then used to update parameters C1 and
R1 , based on the measurements established in Chapter 5. An iterative solution is required to
converge on the appropriate parameters.

6.1 Simulation Techniques

140

Once the volumetric compliance C1 and bulge damping R1 parameters have converged,
the frequency domain results of the excitation X and transmitted force FT are calculated via a
Fourier transform. The cross point dynamic stiffness and phase angle relation are calculated with
the method presented in Section 1.2, and the procedure is repeated for each excitation condition
to determine the complete frequency response curve.
To include the bell model in the above simulation procedure, equations (6.1) and (6.2) are
replaced with new continuity equations, from Section 2.7,

Cb Pb = Qb (Am Ab Ap )X + Cb Rcb Q b (Am Ab Ap )X

(6.16)

C1 P1 = (Am Ab )X Qb Qd Qi QL

(6.17)

C2 P2 = Qi + Qd + QL

(6.18)

The bell momentum equation

Qb = P1 Pb Rb1 (Qb + Ab X) Rb2 Qb + Ab X (Qb + Ab X)


Ib

(6.19)

is added to equations (6.3) and (6.4), to complete the state equations.

Time domain solutions are obtained using the same numerical differentiation technique.
Transmitted force is also updated to include the influence of the bell chamber pressure, from
equation (2.82)
FT = kr X + br X + (Am Ad_f nc )(P1 P2 ) + Am P2
+(Ad Ad_holes ) (Rd1 + Rd2 |Qd |) Qd (Am Ap )Pb
Volumetric compliance parameters Cb (wdr , VT , Fp ) and Rcb (wdr , VT , Fp ), associated with
the bell, are altered according to the volume amplitude entering the chamber.

Total volume

entering the bell chamber is calculated using


VT =

Qb (Am Ap Ab )X dt

(6.20)

6.2 Model Validation

141

The volume amplitude VT is calculated and used to modify Cb and Rcb . An iterative solution is
again applied to obtain volumetric compliance parameters from measured values.
The solution method developed above represents a computationally intensive numerical
differential solution, coupled with an iterative technique to obtain appropriate system parameters. As a result, the solution time becomes an issue. All simulations conducted for this thesis
were run using MATLAB under Windows, on an Intel PII 233MHz processor. Solutions operating in the nonlinear coupled region required up to five minutes of processing time per frequency
solution, representing a significant number of instructions. However, the processing requirements can be significantly reduced by compiling the code into a stand-alone executable code,
eliminating the run time compiling imposed by MATLAB.

6.2

Model Validation

The Cooper-Standard hydraulic mount investigated in Chapter 5 is now assembled and tested
using the standard dynamic characterization procedure described in Section 1.2.

Measured

dynamic stiffness and phase characteristics are compared with the simulated nonlinear model to
validate the modeling and parameter identification approach. The models are simulated using
parameters identified in Chapter 5 and summarized in Table 5.2. Five system tests include: the
inertia track only, the decoupler only, the decoupler and inertia track, the inertia track and bell,
and the inertia track, decoupler and bell.

6.2.1

Inertia Track

In the first case, the response of the inertia track is observed by removing the decoupler from the
mount in Figure 6.2. The simulated model and measured data are compared in Figure 6.3. Both
the phase angle and dynamic stiffness response of the model is within 5% of measured data.
Several conclusions about the modeling technique and parameter identification can be made.

6.2 Model Validation

142

Decoupler

Inertia
Track

Phase (Degrees)

Dynamic Stiffness (N/mm)

Figure 6.2: Illustration of the Cooper-Standard mount being compared with simulation data.

Frequency (Hz)
1mm P-P

2mm P-P

1mm P-P sim

2mm P-P sim

Figure 6.3: Low frequency simulation versus measured response, on a mount with only an
inertia track.

6.2 Model Validation

143

First, the dynamic stiffness and phase characteristics at frequencies beyond the liquid column
resonance demonstrate amplitude dependent behavior. Even without the decoupler, the dynamic
stiffness beyond resonance is higher for the 1 mm P-P (peak-to-peak) amplitude excitation over
the 2 mm P-P response. This nonlinear effect is attributed to the upper chamber rubber stiffness
kr and volumetric compliance C1 parameters. Since the model response captures this effect, the
approach that identifies nonlinear parameters kr , br , C1 and R1 with a local linear response at
each frequency and excitation amplitude, is concluded to be effective. Agreement between the
model and measured response at the resonant frequency, indicates that the inertia track parameter
identification (see Section 5.3) is effective. Also, the phase angle correlation indicates that the
nonlinear resistance model captures the oscillatory fluid damping appropriately.

6.2.2

Decoupler

The Cooper-Standard mount was also assembled and tested with only the decoupler between
chambers. Model simulations under the same conditions are plotted against the measured data
in Figure 6.4.

For both the 0.1 mm P-P and 0.3 mm P-P amplitude conditions, the model

characteristics beyond the liquid column resonant frequency indicate stiffness and compliance
parameters have also been modeled and identified appropriately in the high frequency range.
Perhaps the most significant aspect is the nonlinear decoupler model performance at high
frequencies. Figure 6.4 shows that the amplitude dependent model is within 10% of the measured data. At 140 Hz, a 10% increase in the model damping is attributed to decoupler cage
contact, which introduces a large resistance that halts flow. As explained in Section 3.3, the
decoupler response is pushed into its nonlinear region around resonance at high frequencies.
However, overall the decoupler switching model is effective under high frequency excitations.
Finally, the trends in the 0.1 mm P-P data indicate that the decoupler liquid column resonance is occurring at the correct frequency and is imposing an appropriate amount of damping

144

Phase (Degrees)

Dynamic Stiffness (N/mm)

6.2 Model Validation

Frequency (Hz)
0.1mm P-P

0.3mm P-P

0.1mm P-P sim

0.3mm P-P sim

Figure 6.4: High frequency simulation versus measured response, on a mount with only a
decoupler.
on the system. This shows that the identification process has extracted appropriate inertia and
resistance parameter values for the free decoupler.

6.2.3

Inertia Track and Decoupler

The mount shown in Figure 6.2, with both the inertia track and decoupler, is now characterized
over the complete range of excitations. Figure 6.5 shows an approximate 5% correspondence
between simulated and measured response in the low frequency range. Comparing these results
to Figure 6.3, it is noted that now the 2 mm P-P excitations yield a higher dynamic stiffness
than for the 1 mm P-P excitations. This indicates how the decoupler dominants the amplitude
dependent characteristics, which have been appropriately modeled and identified. These results
also indicate that the decoupler switching model and parameter identification techniques are valid
under low frequency large amplitude excitations.

145

Phase (Degrees)

Dynamic Stiffness (N/mm)

6.2 Model Validation

Frequency (Hz)
1mm P-P

2mm P-P

1mm P-P sim

2mm P-P sim

Phase (Degrees)

Dynamic Stiffness (N/mm)

Figure 6.5: Low frequency simulation versus measured response, on a mount including a decoupler and inertia track.

Frequency (Hz)
0.1mm P-P

0.3mm P-P

0.1mm P-P sim

0.3mm P-P sim

Figure 6.6: High frequency simulation versus measured response, on a mount including a
decoupler and inertia track.

6.2 Model Validation

146

The high frequency system model response in Figure 6.6 is similar to that of a decoupler
alone, and fits the measured data within 5% accuracy. Therefore, the high frequency model with
the decoupler and inertia track is also shown to be effective.

6.2.4

Inertia Track and Bell

The bell system is now included in the Cooper-Standard mount, as illustrated in Figure 6.7. To
begin, only the inertia track and bell are included in the experimental mount to demonstrate
the influence of the bell system. As Figure 6.8 illustrates, the low frequency model response
falls within 10% of the measured data. The peak in dynamic stiffness at 20 Hz corresponds to
the first natural frequency of the inertia track and bell system. This indicates that the bell has
negligible influence on low frequency characteristics; however, the high frequency response is
significantly altered by the bell plate. As Figure 6.9 indicates, the model and measured response
both capture the notch at 200 Hz. Since this system only includes a bell and inertia track, the
notch corresponds to an antiresonance of the inertia track fluid column, similar to that explained
for the decoupler in Section 3.5. In the high frequency range, the model and measured response
are within 5% agreement. These results indicate that the bell model and parameter identification
procedures have accurately captured the system dynamics.

6.2 Model Validation

147

Bell Plate
Bell Position
With Preload

Phase (Degrees)

Dynamic Stiffness (N/mm)

Figure 6.7: Illustration of the Cooper-Standard mount used to validate system models with a
bell.

Frequency (Hz)
1mm P-P

2mm P-P

1mm P-P sim

2mm P-P sim

Figure 6.8: Low frequency simulation versus measured response, on a mount including an
inertia track and bell system.

148

Phase (Degrees)

Dynamic Stiffness (N/mm)

6.2 Model Validation

Frequency (Hz)
0.1mm P-P

0.3mm P-P

0.1mm P-P sim

0.3mm P-P sim

Figure 6.9: High frequency simulation versus measured response, on a mount including an
inertia track and bell system.

6.2 Model Validation

6.2.5

149

Inertia Track, Decoupler and Bell

Finally, the model that includes a bell, decoupler and inertia track is validated. At low frequencies the simulation and experimental data are within approximately 10% deviation, for both the
1 mm P-P and 2 mm P-P excitations. As Figure 6.10 indicates the response is similar to that of
the decoupler and inertia track alone (see Figure 6.5).
The high frequency dynamic characteristics demonstrate the effects of the bell system.
Figure 6.11 shows the amplitude dependent characteristics imposed by the decoupler, as well as
the notch influence of the bell system. However, up to approximately 25% difference between
the model response and measured data is shown. These discrepancies are attributed to a degradation in the upper compliance, occurring in the time between when the parameters were identified
and the bell tests were conducted. The rubber degradation resulted in increased material stiffness, which is demonstrated by the increase in dynamic stiffness at 10 Hz, as compared to those
results shown in Figure 6.6 which were measured several months prior. The decreased compliance has shifted the measured antiresonance frequency to approximately 210 Hz. However, it
is important to focus on how the high frequency trends validate the bell model and parameter
identification techniques.
Figure 6.12 highlights how the bell plate alleviates the high frequency stiffness increase
due to decoupler resonance. Thus, the bell system improves the high frequency characteristics
of a hydraulic mount.

150

Phase (Degrees)

Dynamic Stiffness (N/mm)

6.2 Model Validation

Frequency (Hz)
1mm P-P

2mm P-P

1mm P-P sim

2mm P-P sim

Phase (Degrees)

Dynamic Stiffness (N/mm)

Figure 6.10: Low frequency simulation versus measured response, on a mount including an
inertia track, decoupler and bell system.

Frequency (Hz)
0.1mm P-P

0.3mm P-P

0.1mm P-P sim

0.3mm P-P sim

Figure 6.11: High frequency simulation versus measured response, on a mount including an
inertia track, decoupler and bell system.

151

Phase (Degrees)

Dynamic Stiffness (N/mm)

6.2 Model Validation

Frequency (Hz)
,

without bell plate

with bell plate

Figure 6.12: Comparison of high frequency response with and without the bell plate, using
both measured and modeled response.

6.3 Summary

6.3

152

Summary

This chapter has presented the results of simulation techniques used to obtain frequency domain
model responses with the experimentally identified parameters.

Using the Cooper-Standard

mount, the models are validated with measured response data. The inertia track and decoupler
models, and corresponding parameters, are shown to represent the measured system characteristics. In particular, the decoupler response at high frequencies corresponds with experimental
data and indicates that the nonlinear decoupler model represents the decoupler dynamics.
Small increases in model damping during high frequency cage contact, are attributed to the
resistive flow stopping method imposed by the nonlinear decoupler model. Other discrepancies
are attributed to experimental errors in the measured parameters and general approximations
made with the lumped parameter assumption. However, simulation results presented here are
within acceptable levels by industry standards.
Finally, two mount configurations have been presented that validate the high frequency effects of the bell model. The high frequency notch occurs in both the measured and model
response, thereby validating the modeling approach and explanations regarding the dynamic aspects in Section 3.5.
Having validated the modeling approach and parameter identification procedures, the thesis
will now begin to investigate how some parameters can be associated with mount geometry.

Chapter

Preliminary Studies on Parameter to Geometry Relation


This chapter is focused on providing some insight into the relationship between lumped model
parameters and physical component geometry. Specifically, the work presented here investigates
the inertia track, decoupler and bell parameters, using the experimental procedures described
in Chapter 5. The parameters associated with rubber elements are not considered due to the
complex behavior of the material.

7.1

Inertia Track

To confirm trends regarding the inertia track parameters, three test cases are constructed, as
illustrated in Figure 7.1. The different geometry cases include, (a) an inertia track of average
length and cross section, (b) the same inertia track with screws inserted into the fluid column to
increase resistance, and (c) a long inertia track with a small cross section.

7.1.1

Inertia

The inertia associated with the column of fluid in the inertia track has been established in [5, 15,
18, 30]. To relate the inertia to geometry, the inertia track is considered to have a mass equivalent

153

7.1 Inertia Track

154

L = 212 mm
2
Ai = 57.2 mm

L = 212 mm
2
Ai = 57.2 mm

L = 360 mm
2
Ai = 28.0 mm

Case (a)

Case (b)

Case (c)

Figure 7.1: Illustration of inertia track geometries under investigation.


to the volume of fluid residing in the inertia track. The inertia becomes

Ii =

Mi
L
=
2
Ai
Ai

(7.1)

where L is the inertia track length and Ai represents the cross-sectional area. The fluid density
in this case is 1.028 106 kg/mm3 (1028 1012 Ns2 /mm4 ) composed of 50% water and 50%
ethylene glycol, by volume.
Using the measured length and cross-sectional area, the inertia parameters are calculated
and compared to those measured experimentally.

The data in Table 7.1 confirms that equa-

tion (7.1) can be used to obtain a good approximation of the inertia parameter based on the
geometry.
Table 7.1: Calculated versus measured inertia parameters for the inertia track.

case (a)
case (b)
case (c)

L (mm)
212
212
360

Ai (mm2 )
57.2
57.2
28.0

Measured

Ii kg/mm4
3.29 106
3.84 106
1.32 105

Calculated

Ii kg/mm4
3.81 106
3.81 106
1.31 105

7.2 Decoupler

7.1.2

155

Resistance

Resistance parameters associated with the inertia track are not as well established. Since these
parameters are influenced by such factors as surface roughness, it is difficult to establish a relation
to the geometry. The resistance parameters Ri1 and Ri2 are identified using the test apparatus
and the identification procedure outlined in Section 5.3. Measured parameters for each test case
are included in Table 7.2.
Table 7.2: Measured resistance parameters of the inertia track.

case (a)
case (b)
case (c)

Ri1 kg/s-mm4
6.52 105
1.09 104
6.14 104

Ri2 kg/mm7
4.49 1010
5.30 1010
1.57 109

The total resistance function Rtotal = Ri1 + Ri2 |Qi | is plotted against the inertia track
flow variable in Figure 7.2a. This resistance is also transformed into velocity damping using Ai .
Pressure loss due to resistance changes from

P = Ri1 Qi + Ri2 |Qi | Qi

(7.2)

Ai P = bi1 Vi + bi2 |Vi | Vi

(7.3)

to

where bi1 = Ri1 A2i , and bi2 = Ri2 A3i . Total velocity damping is plotted in Figure 7.2b and
demonstrates that a reasonable approximation may be to select average damping values based
on experimental results, then convert damping to resistance using Ai . It is recommended that
several more tests be conducted to establish quantitative bounds on the inertia track resistance
and damping parameters.

156

Damping (kg/s)

Resistance (kg/s/mm4)

7.2 Decoupler

Flow (mm /s)


Case (a)

Velocity (mm/s)

Case (b)

Case (a)

Case (c)

(a)

Case (b)

Case (c)

(b)

Figure 7.2: Measured resistance parameters for each test case: (a) flow resistance; and (b)
velocity damping.

7.2

Decoupler

The relationship between decoupler parameters and the physical geometry has not been established in any literature reviewed. This study attempts to provide more insight into this relationship by investigating several configurations and their associated parameters.

The inertia

parameter, uncoupled resistance parameters and switching parameters are all considered.

7.2.1

Inertia

The inertia associated with oscillating flow through an orifice is first studied using the two tests
cases illustrated in Figure 7.3.

Two decoupler assemblies without decoupler plates are con-

structed, one with seven 6.3 mm diameter holes and the other with nineteen. Inertia parameters
for each of the two cases are experimentally identified and documented in Table 7.3.
Table 7.3: Effective inertia and fluid column height for a simple orifice.

case (a)
case (b)

Ad holes mm2
218.2
592.3

hf luid (mm)
17.5
20.3

Iorif ice kg/mm4


8.24 108
3.55 108

7.2 Decoupler

Cross Section

157

hfluid

Top View

Ad_holes = 218.2 mm2

Ad_holes = 592.3 mm2

Case (a)

Case (b)

Figure 7.3: Fluid inertia tests of the decoupler orifice with: (a) seven holes; and (b) nineteen
holes.
An interesting property of the measured inertia is the effective fluid column height, denoted
by hf luid . Using the measured orifice inertia Iorif ice and the total area of the decoupler cage holes
Ad_holes , the theoretical fluid column height is calculated with

Iorif ice =

hf luid
Ad_holes

(7.4)

The data summarized in Table 7.3 highlights a curious trend; in both cases the fluid column
height is almost the same. An explanation of this effect involves a detailed look into the fluid
flow behavior and is beyond the scope of this work.
Once the decoupler plate is included, the flow path within the decoupler is altered and the
resulting decoupler inertia changes. Figure 7.4 illustrates how the decoupler plate interrupts
the fluid columns previously investigated. The fluid must now move horizontally within the
decoupler cage to push the decoupler plate. Using this model, a preliminary relationship between
the decoupler features and the effective inertia is developed to be

Id =

(hf luid hgap )


Mplate
+ 2
Ad_holes
Ad_holes

(7.5)

7.2 Decoupler

158

hgap

hfluid

Figure 7.4: Illustration of the decoupler inertia with a decoupler plate.


where Mplate represents the decoupler plate mass and hgap is the decoupler cage height. This
model assumes that the previously investigated fluid columns that are associated with the simple
orifice, still apply.

However, with the decoupler plate in place, the flow inside the cage is

altered such that short columns of fluid are removed from the original columns. The decoupler
plate mass is considered to contribute to the fluid column inertia, signifying the later term of
equation (7.5).
Two additional test cases are developed by placing rubber decoupler plates within the systems previously investigated and shown in Figure 7.3. Case (c) uses the decoupler cage shown
in case (a) with a rubber decoupler plate. Case (d) adds a rubber decoupler plate to case (b).
Table 7.4 lists the physical geometry parameters associated with the two decoupler cases and
compares the calculated inertia parameters to those measured.
Table 7.4: Calculated versus measured inertia for the complete decoupler system.

case (c)
case (d)

A
d holes

mm2
218.2
592.3

hfluid
(mm)
17.5
20.3

hgap
(mm)
4.8
4.8

Mplate
(kg)
2.36 103
8.59 103

Measured
Id 4
kg/mm
9.87 108
4.14 108

Calculated
Id 4
kg/mm
1.09 107
5.14 108

These tests have achieved a preliminary relation between decoupler geometry and the inertia parameter. More research is required to fully understand the fluid flow patterns and effective

7.2 Decoupler

159

inertia. It should be noted, however, that this research has demonstrated the ability to isolate and
study particular components from the hydraulic mount system. The experimental methods have
opened new possibilities for investigating the internal fluid dynamics. For example, it may be
feasible to perform CFD (Computational Fluid Dynamics) analysis on a system representing the
isolated tests. This technique may prove advantageous to hydraulic mount development, since it
would provide insight into how oscillating fluid flow across an orifice generates an effective fluid
inertia.

7.2.2

Resistance

The decoupler resistance parameters are influenced by numerous physical aspects of the decoupler. Fortunately, the decoupler resistance parameters do not dramatically influence the dynamic
characteristics (see Section 3.1). This study does not attempt to determine a relationship between decoupler geometry and resistance parameters, however resistance data is plotted to highlight trends. Table 7.5 indicates the experimentally measured parameters for each of the test
cases considered in the inertia studies. Figure 7.5 includes plots of the resistance parameters
versus the decoupler flow. The measured results show how the decoupler plate increases the resistance, as flow is forced to change direction inside the cage. Also, cases (c) and (d) indicate a
moderate decrease in resistance as the number of holes increases.
Table 7.5: Measured decoupler resistance parameters.

case (a)
case (b)
case (c)
case (d)

Rd1 kg/s-mm4
1.29 106
5.57 107
4.10 106
3.13 106

Rd2 kg/mm7
2.35 1011
6.51 1012
5.45 1011
4.27 1011

160

Resistance (kg/s/mm4)

7.2 Decoupler

Flow (mm3/s)
Case (a)

Case (b)

Case (c)

Case (d)

Figure 7.5: Measured decoupler resistance parameters plotted for each of the decoupler cases
investigated.

7.2.3

Switching

The final decoupler properties to be investigated are the switching parameters Rd3 and Vd_ max .
These parameters are used in the nonlinear model to limit fluid volume through the decoupler, and
they are also used in the functions Ad_f nc and RL2_f nc . As stated in Chapter 3, these parameters
are related through the decoupler area Ad . To establish this relationship, experimental data is
collected from several decoupler tests, using the identification procedure described in Section 5.4,
and plotted in Figure 7.6. The trendline equation is
Vd_ max
= 15.5 (Rd3 )0.9739
Ad

(7.6)

To obtain the Rd3 parameter given the decoupler geometry, the maximum volume of fluid Vd_ max
is calculated using

Vd_ max =

d
Ad
2

(7.7)

161

Vd_max / Ad (mm)

7.3 Bell Parameters

Parameter Rd3
Measured

Trendline

Figure 7.6: Measured decoupler displacement versus the decoupler switching parameter.
where d denotes the available displacement of the decoupler plate, calculated with

d = hgap hplate

(7.8)

The parameter hplate represents the thickness of the decoupler plate shown in Figure 7.7.
Table 7.6 includes a comparison of the volume amplitude parameters measured from experimental data and calculated from geometry. Comparing the calculated and measured Vd_ max
parameters for cases (c) and (d), it appears as though the experimental identification has picked
up on a thin layer of fluid that slightly reduces the decoupler travel. If 0.5 mm is removed from
hgap the calculated and measured data show improved correlation.

Since these studies have

provided only preliminary results, further studies should be conducted to validate the decoupler
parameters.
Table 7.6: Calculated decoupler switching parameters versus measured.

case (c)
case (d)

Ad 2
mm
945
3216

hgap
(mm)
4.8
4.8

hplate
(mm)
2.6
2.6

Measured
Rd3
18.2
19.0

Measured
Vd max

mm3
867
2707

Calculated
Vd max

mm3
1040
3537

7.3 Bell Parameters

162

hplate
hgap
Ad
Figure 7.7: Illustration of decoupler geometry used to calculate the switching parameters.

7.3

Bell Parameters

The bell parameters were not studied due to time constraints, however a few comments can
be made regarding the inertia parameter shown in Table 7.7.

First, the fluid column height,

associated with the bell, can be estimated using

Ib =

hf luid
Ab

(7.9)

The bell fluid column height is found to be within the range of both decoupler tests.

It is

recommended that more effort be placed into investigating this fluid column, similar to the further
studies recommended for the decoupler inertia.
The second point to note is the relatively low resistance parameters associated with the bell
fluid column. As mentioned in Section 3.5, the bell resistance parameters do not have significant
impact on the dynamic characteristics, thus it is not recommended that extensive effort be placed
on developing analytical equations to relate resistance parameters to the geometry.
Table 7.7: Measured bell parameter, indicating fluid column height.

measured

Ab 2
mm
600

Ib 4
kg/mm
2.24 108

Rb1 4
kg/s-mm
2.412 107

Rb2 7
kg/mm
3.74 1012

hf luid
(mm)
13.1

7.4 Summary

7.4

163

Summary

This chapter has provided some preliminary insight into how the inertia track, decoupler and bell
parameters relate to their respective geometry. First, a confirmation has been made regarding
the relation between inertia track geometry and the inertia parameter. The resistance parameters
pertaining to the inertia track are documented, however no conclusive relationship with geometry has been made. Next, the decoupler was investigated. Experimental results for the inertia of
oscillating flow across an orifice were presented and used to establish an effective fluid column
height. A preliminary model for the complete decoupler was developed. The decoupler switching parameters have also been written in terms of the geometry and compared with measured
results. Finally, the bell parameters are documented, but no formal study was conducted.
This chapter has proposed several areas for further research. It is suggested that the inertia
track flow resistance should be converted to velocity damping to establish reasonable relationships with the geometric features. More isolated tests are required to determine the trends.
Resistance to flow across the free decoupler and bell do not have significant impact upon
the dynamic characteristics, thus it is not recommended that extensive effort be placed on developing analytical equations to relate these resistance parameters to the geometry.
The effective fluid column inertia that is associated with oscillating flow through an orifice
needs to be fully understood. Since this thesis has established an effective experimental procedure for identifying parameters, it is now recommended that tools such as CFD be applied to
appreciate the fluid behavior. These techniques will also help to understand the bell fluid column
inertia.

Chapter

Conclusions and Recommendations


The main goal of this thesis was to develop modeling information that can contribute to reducing the hydraulic mount design period. Development time can be significantly reduced with a
reliable fluid mount model that can be simulated to predict the system response before it is physically assembled. Furthermore, it is important for designers to understand how fluid column
dynamics within the hydraulic mount alter the dynamic stiffness and phase characteristics.
The work presented in this thesis builds on recently published hydraulic mount models,
by developing a nonlinear decoupler model that is effective over the complete frequency range
of interest. Further contributions include a model that captures the effect of adding a bell plate
within the upper chamber.
Dynamic stiffness and phase characteristics of the models are explained in terms of the
internal fluid column dynamics.

These explanations include the standard linear model, the

nonlinear decoupler model, and the linear model that includes the bell plate system.
This thesis also presented a unique experimental approach to identifying hydraulic mount
model parameters. The experimental configuration and identification procedures are presented
to enable an extension of this work to encompass other mount designs. Validation of the models
was accomplished by simulating the equations using experimentally identified parameters and
comparing the results with measured mount data.

164

8.1

Conclusions

165

In conclusion, the models presented within this thesis will provide an excellent tool for
hydraulic mount design. Also, the explanations will give engineers a better understanding of the
internal fluid column dynamics.

8.1

Conclusions

Several conclusions are made that pertain to certain aspects of this work.

The nonlinear decoupler model presented in this work uses flow across the decoupler
in combination with volume and, as a result, is effective over the complete frequency
spectrum of interest. This approach is more effective at high frequencies, since the
decoupler inertia will force the fluid column motion out-of-phase with the pressure
gradient.

A lumped parameter bell model has been established by associating an inertia with the
annular column of fluid between the bell plate and mount wall. Dynamic characteristics
of the bell and decoupler system have been physically explained using a linear, high
frequency, two DOF model. Modal analysis techniques have associated the high
frequency notch, indicative of a bell system, with an antiresonance of the decoupler
fluid column. This high frequency decrease in dynamic stiffness is a function of the bell
chamber compliance, fluid column inertia and effective area parameters within the mount.

Other enhancements to the model include the addition of bulge damping to the upper
chamber. This is accomplished with a resistance parameter on the fluid accommodated by
the volumetric expansion. The parameter is identified and validated in the system models.

8.1

Conclusions

166

Each of the upper chamber compliance parameters were identified through one of three
tests. The stiffness and damping parameters were isolated by testing a dry mount, and
the effective pumping area was isolated in the testing fixture and plotted as a function
of mount displacement. The upper chamber volumetric compliance parameters were
extracted using a mount with a long narrow inertia track and no decoupler. Lower
chamber volumetric compliance characteristics were extracted using the test apparatus
and are only a function of preload conditions.

The inertia and resistance parameters pertaining to the inertia track were best extracted
using the least-squares method with random flow perturbations. However, the decoupler
identification procedure required two separate testing procedures. First, a frequency
sweep test was conducted to establish the uncoupled inertia and resistance parameters.
Large amplitude random excitations were then used with the Extended Kalman Filter
to extract the switching parameter and leak flow resistance. The liquid column inertia
and resistance associated with the bell were also identified using frequency sweep
measurements.

From the system results it is concluded that the approach of identifying nonlinear
parameters kr , br , C1 and R1 with a local linear response at each frequency and excitation
amplitude, is effective.

Agreement between the modeled and measured response also have determined that the
inertia track model and parameter identification are effective. Damping correlation
indicates that the nonlinear resistance model captures the oscillatory fluid damping
appropriately.

8.2 Recommendations

167

At high frequencies, the decoupler response matches experimental data and indicates that
the nonlinear decoupler model represents the decoupler dynamics. Results show how the
amplitude dependent model enters its nonlinear region during larger motions, typically
occurring around resonance at high frequencies. Further results also indicate that the
decoupler switching model and parameter identification techniques are valid under low
frequency large amplitude excitations.

Two mount configurations have been presented that validate the high frequency effects
of the bell model. The high frequency notch occurs in both the measured and model
response, thereby indicating that the modeling approach and the explanations regarding
the dynamic behavior are accurate.

8.2

Recommendations

Recommendations for future research are made regarding the initial work to establish parameterto-geometry relationships.

It is suggested that the inertia track flow resistance parameters can be converted to velocity
damping parameters for a better association with the geometric features. However, more
isolated tests are required to fully establish these trends. Several tests should be conducted
to establish the relationship between damping parameters and general features such as
surface roughness in the internal channel.

The effective fluid column inertia that is associated with oscillating flow through an orifice
needs to be fully understood. Since this thesis has established an effective experimental

8.2 Recommendations

168

procedure, it is now recommended that tools such as CFD be applied to fully appreciate
the fluid dynamic behavior. A CFD model can be constructed to represent the two
chamber test apparatus and validated with measured data. The computational flow
patterns and pressure regions will help to gain a complete appreciation for the lumped
fluid column inertia.

169

Bibliography
[1]

M. Bernuchon. A new generation of engine mounts. SAE Technical Paper Series 840259,
1984.

[2]

P. P. J. V. D. Bosh and A. V. der Klauw. Modeling, Identification and Simulation of


Dynamical Systems. CRC Press Inc., Boca Raton, Florida, 1994.

[3]

J. Colgate, T. Chang, C. Chiou, K. Liu, and M. Kerr. Modelling of a hydraulic engine


mount focusing on response to sinusoidal and composite excitations. Journal of Sound
and Vibration, 184:503528, 1995.

[4]

P. Corcoran and G. Ticks. Hydraulic engine mount characteristics. SAE Technical Paper
Series 840407, 1984.

[5]

W. Flower. Understanding hydraulic mounts for improved vehicle noise, vibration and
ride qualities. SAE Technical Paper Series 850975, 1985.

[6]

S. Gau and J. Cotton. Experimental study and modeling of hydraulic mount and engine
system. SAE Technical Paper Series 951348, 1995.

[7]

D. J. Inman. Engineering Vibration. Prentice Hall, Toronto, ON, 1996.

[8]

M. Ishihama, S. Satoh, K. Seto, and A. Nagamatsu. Vehicle vibration reduction by transfer function phase control on hydraulic engine mounts. JSME International Journal Series C,, 37(3):536541, 1994.

[9]

J.-N. Juang. Applied System Identification. Prentice-Hall Canada Inc., Toronto, 1994.

[10] G. Kim. Study of Passive and Adaptive Hydraulic Mounts. PhD thesis, Ohio State University, 1992.
[11] G. Kim and R. Singh. Resonance, isolation and shock control characteristics of automotive nonlinear hydraulic engine mounts. Transportation Systems, proceedings of
ASME/WAM DSC, 44:165180, 1992.
[12] G. Kim and R. Singh. Non-linear analysis of automotive hydraulic engine mounts. ASME
Journal of Dynamic System Measurement and Control, 115:482487, 1993.
[13] G. Kim and R. Singh. A study of passive and adaptive hydraulic engine mount systems
with emphasis on non-linear characteristics. Journal of Sound and Vibration, 179:427
453, 1995.

170

[14] Lawrence, F. Shampine, and M. Reichelt. The MATLAB ODE suite. Society for Industrial and Applied Mathematics, 18:122, 1997.
[15] K. Lee and Y. Choi. Performance analysis of hydraulic engine mount by using bond
graph method. SAE Technical Paper Series 951347, 1995.
[16] E. Little and R. Kashani. Adaptive-passive and active hydraulic engine mounts. ASME
Advanced Automotive Technologies, DSC-Vol.56/DE-Vol.86:135140, 1995.
[17] Maia, Silva, He, Lieven, Lin, Skingle, To, and Urgueira. Theoretical and Experimental
Modal Analysis. John Wiley and Sons Inc., Toronto, ON, 1997.
[18] D. Margolis and R. Wilson. Modeling, simulation and physical understanding of hydromounts. Masters thesis, University of California, 1997.
[19] The Math Works, Inc., 24 Prime Park Way, Natick MA. MATLAB 5.2, 1998.
[20] R. Matthew and A. Haddow. On the dynamic response of hydraulic engine mounts. SAE
Technical Paper Series 931321, 1993.
[21] MTS Systems Corporation, Eden Prairie, MN 55344-9764. TestStar Installation Manual,
1996.
[22] MTS Systems Corporation, Eden Prairie, MN 55344-9764. TestStar Dynamic Characterization, 1996.
[23] M. Muller, H. Eckel, M. Leibach, and W. Bors. Reduction of noise and vibration by
approximate engine mount systems and active absorbers. SAE Technical Paper Series
960185, 1996.
[24] M. Muller, O. Weltin, C. Freudenberg, D. Law, M. Roberts, and T. Siebler. Engine
mounts and NVH. Automotive Engineering / July, 1994.
[25] D. Rowell and D. Wormley. System Dynamics: An Introduction. Prentice-Hall, Inc.,
Toronto, 1997.
[26] T. Royston and R. Singh. Periodic response of non-linear engine mounting systems. SAE
Technical Paper Series 951297, 1995.
[27] T. Royston and R. Singh. Optimization of passive and active non-linear vibration mounting systems based on vibratory power transmision. Journal of Sound and Vibration,
194(3):295316, 1996.

171

[28] T. Royston and R. Singh. Study of nonlinear hydraulic engine mounts focusing on decoupler modeling and design. SAE Technical Paper Series 971936, 1997.
[29] K. Seto, K. Sawajari, A. Nafamatsu, M. Ishihama, and K. Doi. Optimum design method
for hydraulic engine mounts. SAE Technical Paper Series 911055, 1991.
[30] R. Singh, G. Kim, and V. Ravindra. Linear analysis of automotive hydo-mechanical
mount with emphasis on decoupler characteristics. Journal of Sound and Vibration,
158(2):219243, 1992.
[31] T. Ushijima and T. Dan. Nonlinear b.b.a. for predicting vibration of vehicle with hydraulic
engine mounts. SAE Technical Paper Series 860550, 1986.
[32] T. Ushijima, K. Takano, and H. Kojima. High performance hydraulic mount for improving vehicle noise and vibration. SAE Technical Paper Series 880073, 1988.
[33] F. M. White. Fluid Mechanics. McGraw-Hill, Inc., Toronto, 1994.
[34] W. Wilson. System identification course notes. 1999.

Appendix

Bond Graph Modeling


The basic idea behind the bondgraph formulation is to keep track of the flow of power within a
physical system. This is accomplished by separating a physical system into components or subsystems. Components in bondgraphs are categorized as 1-port, 2-port and multi-port elements.
It should be noted that there is a difference between the number of physical connections and the
number of ports of an element. A simple example of this is the spring. A spring is connected
between two masses however it can only exchange energy between the spring net displacement
and its force. Thus it is a 1-port element.
In bondgraph, we use two generalized variables for any dynamic systems. These variables are
flow and effort denoted by e(t) and f(t), respectively. The multiplication of the variables in
each system should represent the power flowing between elements. For this reason, the flow and
effort are called power variables. For example, in mechanical systems velocity and force and in
electrical systems current and voltage are selected as the flow and effort variables.
Along with the power variables, there are two other generalized variables called momentum, p(t),
and displacement, q(t), which are called energy variables. The momentum p(t) and displacement
q(t) are defined as the time integral of an effort and flow as

172

Appendix A Bond Graph Modeling

p(t) =

e(t)dt

q(t) =

173

f (t)dt
0

Using the energy variables, the power variables can be written as follows:
e(t) = p(t)

f (t) = q(t)

As mentioned earlier, bondgraph elements may have one, two or multiple ports. 1-port elements
include resistance or R-elements, inertia or I-elements and capacity or C-elements. In addition,
there are ideal effort and flow sources that are considered as 1-port elements.
The R-elements generalize the electrical resistance and have a functional relation between the
effort and flow variables as: e = gR (f ). The I-elements generalize the mechanical inertia in the
bondgraph formulation. I-elements have a functional relation between the flow and momentum
variables as: f = gI (p). The C-elements generalize the mechanical springs or electrical capacitors. In this element, there is a functional relation between the effort and displacement variables
as: e = gC (q). Table A.1 shows the functional relation and symboles used for each element in
bondgraph.
Ideal power sources in bondgraphs are ideal effort and flow sources. An ideal effort source is
an element whose effort is only a function of time and independent of its flow. An ideal flow
source is an element whose flow is only a function of time and independent of its effort. Table
A.1 shows the symbols of ideal sources in bondgraphs. It should be noted that in bondgraphs
is used to indicate energy flow while is reserved to indicate information flow.
Transducers or 2-port elements in the bondgraph formulation can exchange energy through two
ports. Electrical motors and hydraulic pumps are examples of transducers. In an ideal transducer
(see TableA.1), there is no loss and the input/output relation is static and linear. The lossless
assumption implies that e1 f1 = e2 f2 and the static and linear assumptions imply that only two
types of tranducers can exist: 1) Transformers where the effort/flow relationships are: e1 = me2 ,

Appendix A Bond Graph Modeling

174

Table A.1: Bondgraph Elements


Element

Functional Relation

Symbol

R-Element

e = gR (f)

e
f

I-Element

f = gI (p)

e
f

C-Element

e = gC (q)

e
f

Ideal Flow Source

Sf

Ideal Eort Source

Se

Transducer
Transformer
Gyrator

e1

e1 f1 = e2 f2
e1 = me2 , f1 =

f1

1
m f2

e1 = rf2 , f1 = 1r e2

e1
f1
e1
f1

e1

0-Junction

e1 = e2 = e3 , f1 + f2 f3 = 0

f1
e1

1-Junction

f1 =

1
f
m 2

f1 = f2 = f3 , e1 + e2 e3 = 0

f1

e2

TPT

f2
e2

TF

f2
e2

GY

f2

e3

f3
e2 f2
1

e3
f3

e2 f2

and 2) Gyrators where the effort/flow relationships are: e1 = rf2 , f1 = 1r e2 where

m and r are the transformer and gyrator ratio. In a transformer, an effort (flow) variable is
proportional to an effort (flow) variable via the transformer ratio, while in a gyrator an effort
(flow) variable is proportional to a flow (effort) variable via the gyrator ratio. Examples of
transformers are rigid and massless levers, electrical transformers, gear trains and hydraulic rams.
An example of a gyrator is the ideal electric direct current motor, where the torque is proportional
to the current and the armature voltage is proportional to the rotational speed.
The final bondgraph elements are the multi-port elements or junctions. Similar to transducers,
energy is conserved in junctions. There are only two types of junctions: 1) the 0-junction and, 2)
the 1-junction. In a 0-junction the efforts are equal and the algebraic sum of the flows is always
zero. In a 1-junction on the other hand, the flows are equal and the algebraic sum of the efforts is
always zero. The 1-junction in electrical systems represents Kirchhoffs voltage law written for

Appendix A Bond Graph Modeling

175

a loop. The 1-junction for an inertial element is the direct implementation of Newtons 2nd law.
Table A.1 shows the functional relation and symboles used for 0- and 1-junctions.
The final topic of discussion that needs to be addressed is causality. As discussed so far, all
elements exchange energy through their effort and flow variables. In addition, there is a constitutive equation which is the mathematical description of the element and may be dynamic or
static. Because of the constitutive equation one variable is the input or cause and the other is the
output or effect of the element. Since the variables with which any element is interacting in a
bondgraph model are effort and flow, there are two ways of describing an element behaviour, 1)
effort input and flow output or, 2) flow input and effort output. These two situations are called
the causal forms of an element. The graphical symbol used to show causality is a straight vertical line (causal stroke) drawn at one of the ends of energy bond:

or

. The element

nearest the causal stroke has effort impressed on it as input and produces flow as output. Similarly, the element at the other of the bond has flow imposed on it as input and produces effort as
output.
Table A.2: Causality of Bondgraph Elements
Element

Causality
R

R-Element

I-Element

C-Element

Ideal Flow Source

Sf

Ideal Eort Source

Se

Transformer
Gyrator
0-Junction
1-Junction

TF
TF

GY
GY

Appendix A Bond Graph Modeling

176

The causality of sources depend on whether they are a source of effort or flow. In R-elements,
where an effort is related to a flow with an algebraic relation, there is no preference in causality selection as long as the effort or flow can be expressed as a function of each other uniquely.
However, in C- and I-elements the preferred causality is when a flow and effort are the inputs,
respectively. The preferred causality for a C-element is:

and for an I-element is

This choice of causality is called an integral causality and the outputs (flow and effort) are obtained using integration and not differentiation. Table A.2 shows preferred causality for different
bondgraph elements.
Using the functional equations of transducers shown in Table A.1, the casaulity can be defined.
There are two possible causalities for transformers and gyrators which are shown in Table A.2.
In the 1-junction the flow in all ports is the same and therefore, more than one flow cannot be
defined as being the input. The flow of the other ports will be the outputs of the junction. In
the 0-junction the effort is the same in all ports and therefore, more than one effort cannot be
defined as the input. The effort of the other ports will be the outputs of the junction. Table A.2
summarizes the causality discussion of bondgraph elements.
To this end, we have shown that the modeling of a dynamic system in the bondgraph method
is based upon; 1) identifying all components, 2) connecting them to each other through energy
bonds and 0- and 1-junctions, and 3) assigning causality to each bond using preferred causality.
The last stage in bondgraph modeling it to extract the state-space equations of the system, which
can be obtained by using the following procedure:

Choosing state variables. Each energy storing element (an I- or C-element) with integral
causality has independent initial conditions which gives rise to one state variable. The
number of state variables of a system is equal to the number of energy storing elements
with integral causality. We select the momentum p and displacement q variables of all

Appendix A Bond Graph Modeling

177

independent energy storing elements to produce the state vector. The input vector includes
all sources in the system.
Write state equations. Reading from the causal strokes on the bondgraph, write the
constitutive equations of the independent energy storing elements. From the junction that
each independent energy storing element is connected to, write an expression for p and q.

Substitute extra variables. In the p and q expressions, substitute all non-state variables
using the constitutive equations and sources, considering the causal strokes on the
bondgraph.

Appendix

Frequency Domain Equation Development


This appendix contains the mathematical development of the frequency domain equations from
the models presented in Chapter 2. The frequency domain equations are used for linear model
simulations presented in Chapter 3. Specifically, the derivations covered here include:

The linear inertia track model

The bell and decoupler system

178

Appendix B Frequency Domain Equation Development

B..1

179

Linear Inertia Track Model

The inertia track model is first considered as a linear, large amplitude, low frequency model. The
internal dynamic system is characterized by equations (2.11) to (2.13), repeated here,
C1 P1 = Ap X Qi

(B.1)

C2 P2 = Qi

(B.2)

P1 P2 = Ii Q i + Ri Qi

(B.3)

and the transmitted force to the mount base, from equation (2.14) is,
FT = kr X + br X + Ap (P1 P2 ) + Ap P2

(B.4)

To manipulate the model into a single frequency domain equation, each system variable is converted using the steady-state sinusoidal solution in exponential form:
P1 = P1 ejwdr t
P2 = P2 ejwdr t

(B.5)

i ejwdr t
Qi = Q
jwdr t
X = Xe
Using these transformations, equations (B.1) and (B.2) become
Ap
1
P1 =
X +j
Qi
C1
wdr C1
1
P2 = j
Qi
wdr C2

(B.6)
(B.7)

Substituting (B.6) and (B.7) into equation (B.3) gives,


Ap
X=
C1

Ri + j Ii wdr

1
1

wdr C1 wdr C2

i
Q

(B.8)

Appendix B Frequency Domain Equation Development

180

i gives
and solving for Q
i =
Q

C2 Ap wdr

X
2
Ri C1 C2 wdr + j (Ii C1 C2 wdr
C2 C1 )

(B.9)

The denominator of equation (B.9) represents the characteristic equation of the single degree-offreedom system and can be shown to posses an eigenvalue of
s
(1/C1 + 1/C2 )
wdr =
Ii

(B.10)

if damping is neglected.

Using (B.9) each of the pressure variables can be written in terms of the input amplitude X
P1
P2

Ap
C2 Ap j
=
+
2
2
C
Ri C1 C2 wdr + j (Ii C12 C2 wdr
C2 C1 C12 )
1

Ap j

X
=
2
Ri C1 C2 wdr + j (Ii C1 C2 wdr
C2 C1 )

(B.11)
(B.12)

To obtain the dynamic stiffness characteristics, equations (B.11) and (B.12) are substituted into
the transmitted force (B.4) and then divided by the input amplitude
K (wdr j) =

A2p
C2 A2p j
Ft
= kr + br wdr j +
+
2

C1 Ri C12 C2 wdr + j (Ii C12 C2 wdr


C2 C1 C12 )
X

(B.13)

The magnitude and phase angle of the complex equation (B.13) are used to obtain the dynamic
stiffness and phase angle, as follows
K = |K (wdr j)|
= ]K (wdr j)

B..2

Bell and Decoupler System

The following section presents the development of frequency domain equations for the bell and
decoupler high frequency linear model using modal analysis. The high frequency linear model

Appendix B Frequency Domain Equation Development

181

that includes the decoupler and bell dynamics, from equations (3.19) to (3.23), are:
Cb P b = Qb (Am Ab Ap )X

(B.14)

C1 P1 = (Am Ab )X Qb Qd

(B.15)

C2 P2 = Qd

(B.16)

P1 Pb = Ib Q b + Rb (Qb + Ab X)

(B.17)

P1 P2 = Id Q d + Rd Qd

(B.18)

The transmitted force to the base, from equation (3.24), is

FT = kr X + br X + (Am Ad )(P1 P2 ) + Am P2 + Ad Rd Qd (Am Ap )Pb

(B.19)

Equations (B.14) to (B.16) are then substituted into equations (B.17) and (B.18) to yield the two
coupled dynamic equations

1
1
1
A

A
A

A
m
b
m
p
b
b + Rb Q b +

+
Qd =
+
Ib Q
Qb +
X Rb Ab X
C1 Cb
C1
C1
Cb

1
1
1
Am Ab

Id Qd + Rd Qd +
Qb +
+
Qd =
X
(B.20)
C1
C1 C2
C1
These equations are converted to matrix form

Ib 0
0 Id

b
Q
d
Q

Rb 0
0 Rd

Q b
Q d
"

Am Ab
C1

1
C1

+
1
C1

1
Cb

1
C1

Am Ap Ab
Cb
Am Ab
C1

1
C1

+
#

1
C2

Qb
Qd

Rb Ab
0

(B.21)

The steady-state sinusoidal solution is expressed in the exponential notation:


b ejwdr t
Qb = Q
d ejwdr t
Qd = Q
jwdr t
X = Xe

(B.22)

Appendix B Frequency Domain Equation Development


Substituting these solutions into equation (B.21) yields

1
2
Ib wdr
+ Rb wdr j + C11 + C1b
C1
1
2
Id wdr
+ Rd wdr j +
C1

1
C1

1
C2

182

b
Q

Qd

Fb
Fd

(B.23)

where,

Fb
Fd

"

Am Ab
C1

Am Ap Ab
wdr j
Cb
Am Ab
C1

2
+ Rb Ab wdr

is used to simplify the terms.


Inverting the coupled matrix (B.23) to solve for the bell and decoupler flow gives,

2
Id wdr
+ Rd wdr j + C11 + C12
C11
Fb

2
b
C11
Ib wdr
+ Rb wdr j + C11 + C1b
Fd
Q
(B.24)
=
d
Q
characteristic equation
where the Characteristic Equation is the determinant of the 2 2 matrix in equation B.23 and
given by

2
det K M wdr
+ Cwdr j =

Ib
Ib
Id
Id
4
2
Id Ib wdr
+
+
+
+ Rb Rd wdr
C1 C2 C1 Cb

1
1
1
+
+
+
C1 C2 C1 Cb Cb C2

3
+ (Ib Rd Rb Id ) wdr
(B.25)

Rb Rb Rd Rd
+
+
+
+
wdr j
C1 C2 C1
Cb

The system matrices M , K and C are taken from equation (B.21)

1
+
Rb 0
Ib 0
M=
; C=
; K = C1 1
0 Id
0 Rd
C1
Solving for the bell and decoupler flow yields

2
Rb Ab wdr
(Am Ab )wdr
1
1
2
I
w
+
+
+
R
w
j
+
d
d
dr
dr
C1
C1
C2
(Am Ab )
b =
Q
Characteristic Equation

1
Cb

1
C1

1
C1

1
C2

(Am Ap Ab )
(Am Ab )Cb

+1 j

1
C1

(B.26)

d =
Q

(Am Ab )wdr
C1

1
Cb

Am Ap Ab
(Am Ab )Cb

2
Ib wdr

+ Rb

Characteristic Equation

Rb Ab
Am Ab

wdr j

(B.27)

Appendix B Frequency Domain Equation Development

183

Finally, the cross point dynamic stiffness response function is determined by writing equations
(B.14) to (B.16) in the frequency domain
1
(Am Ab Ap )
Qb +
X
wdr Cb
Cb
Am Ab
1
1
=
X +j
Qb + j
Qd
C1
wdr C1
wdr C1
1
= j
Qd
wdr C2

Pb = j

(B.28)

P1

(B.29)

P2

(B.30)

The dynamic stiffness is then developed from the transmitted force (B.19) in the frequency domain as

1
Ft
d (Am Ap )Pb (B.31)
= kr + br wdr j +
(Am Ad )(P1 P2 ) + Am P2 + Ad Rd Q

X
X

The magnitude and phase angle of the complex equation (B.31) are used to obtain the dynamic
stiffness and phase angle.

Appendix

Detailed Drawings of Experimental Apparatus


The following Appendix Includes:

Bill of Materials

Assembly Drawings

Detailed Drawings of Each Component

184

Appendix C Detailed Drawings of Experimental Apparatus

185

Appendix C Detailed Drawings of Experimental Apparatus

186

Appendix C Detailed Drawings of Experimental Apparatus

187

Appendix C Detailed Drawings of Experimental Apparatus

188

Appendix C Detailed Drawings of Experimental Apparatus

189

Appendix C Detailed Drawings of Experimental Apparatus

190

Appendix C Detailed Drawings of Experimental Apparatus

191

Appendix C Detailed Drawings of Experimental Apparatus

192

Appendix C Detailed Drawings of Experimental Apparatus

193

Appendix C Detailed Drawings of Experimental Apparatus

194

Appendix C Detailed Drawings of Experimental Apparatus

195

Appendix C Detailed Drawings of Experimental Apparatus

196

Appendix C Detailed Drawings of Experimental Apparatus

197

Appendix C Detailed Drawings of Experimental Apparatus

198

Appendix C Detailed Drawings of Experimental Apparatus

199

Appendix C Detailed Drawings of Experimental Apparatus

200

Appendix C Detailed Drawings of Experimental Apparatus

201

Appendix C Detailed Drawings of Experimental Apparatus

202

Appendix

The Extended Kalman Filter


The Extended Kalman Filter algorithm [9, 34] is presented in the following. The algorithm is
separated into five main operations, starting with the nonlinear system representation,
k+1 = f (X
k , Uk , k) + wk
X
k , Uk , k) + qk
Zk = g(X
k and measurement noise qk on the output Zk .
which includes disturbance noise wk on the state X

1.

From the current value of the states at position k, predict the next state values using the state
equations and the system input Uk
pre = f (X
k , Uk , k)
X
k+1
The state vector notation is expanded to clarify the procedure
pre
X1,k+1
f1 (X1,k , X2,k ...Xn,k , Uk , k)
X pre f2 (X1,k , X2,k ...Xn,k , Uk , k)
2,k+1
pre =
X

=
..
..
k+1


.
.
pre
fn (X1,k , X2,k ...Xn,k , Uk , k)
Xn,k+1
For the first iteration, the initial value of the states must be given.

2.

The output covariance is also predicted using


pre
Ck+1
= Fk Ck FkT + wk

203

Appendix D The Extended Kalman Filter

204

where Ck is the current n n diagonal covariance matrix, or the initial covariance matrix
value for the first iteration. Matrix Fk is a Jacobian matrix of the state equations, evaluated
k
at the current value of the states X

Fk =

f1
X1
f2
X1

f1
X2
f2
X2

..
.

...
...
..
.

f1
Xn
f2
Xn

fn
X1

fn
X2

...

fn
Xn

..
.

..
.

The disturbance noise wk is a constant n n diagonal matrix representing gaussian noise


covariance on each state variable.
3.

The Kalman Gain K is next calculated using the predicted covariance matrix

pre
pre T 1
K = Ck+1
GTk qk + Gk Ck+1
Gk
k , Uk , k), evaluated using
The matrix Gk represents the Jacobian of the output function g(X
k
the current value of the states X
Gk =

g
X1

g
X2

...

g
Xn

The constant qk is a scalar value representing the covariance on the output.


4.

Using the Kalman Gain, the predicted state values are updated to optimized estimates

k+1 = X
pre + K Zk g(X
pre , Uk , k)
X
k+1
k+1
The Kalman gain acts on the difference between the measured output Zk and the calculated
pre .
output using the function evaluated at the predicted next state values X
k+1

5.

Finally, the predicted covariance matrix is updated using the Kalman Gain
pre
pre
Ck = Ck+1
KGk Ck+1

Appendix D The Extended Kalman Filter

205

Steps 1 through 5 are repeated for each input and output sample, solving for the best estimate of
k . For each iteration the Jacobian matrices Fk and Gk must be recalculated for
the state matrix X
proper local system linearization.

Вам также может понравиться