Sie sind auf Seite 1von 175

Null-E Magnetic Bearings

A Dissertation
Presented to
the Faculty of the School of Engineering and Applied Science
University of Virginia

In Partial Fulfillment
of the Requirements for the Degree

Doctor of Philosophy
(Mechanical and Aerospace Engineering)

by

Alexei Filatov
August 2002

Abstract
Using electromagnetic forces to suspend rotating objects (rotors) without mechanical contact is often an appealing technical solution. However, in real life magnetic suspensions have to satisfy many
engineering performance requirements beyond the simple compensation for the rotor weight. These
typically include adequate load capacity and stiffness, low rotational loss, low price, high reliability
and manufacturability. With recent advances in permanent-magnet materials, the magnitudes of the
required forces can often be obtained by simply using the interaction between permanent magnets.
While a magnetic bearing based entirely on permanent magnets could be expected to be inexpensive,
reliable and easy to manufacture, a fundamental physical principle known as Earnshaws theorem
maintains that this type of suspension cannot be statically stable. Therefore, some other physical
mechanisms must be included.
One such mechanism employs the interaction between a conductor and a non-uniform magnetic field in relative motion. Its advantages include simplicity, reliability, wide range of operating
temperature and system autonomy (no external wiring and power supplies are required). The disadvantages of the earlier embodiments were high rotational loss, low stiffness and load capacity.
It was realized, however, that rotational loss, load capacity and stiffness depend strongly on the
topology of the conductors and the magnetic fields.
In theory, the rotational loss in the equilibrium position in the absence of external loading can
be made zero by designing a system such that no electric field develops in the conductor during the
rotor rotation. In this dissertation, we introduce the term Null-E to describe this condition. In the
earlier embodiments, the Null-E condition could not be satisfied exactly.
Load capacity and stiffness can be also maximized through choosing shapes of the conductors
and the fields. From this point of view the field and the conductor shapes used in the so called
Null-Flux Bearings were found to be advantageous: the conductors were shaped as planar loops
with central openings and the fields were orthogonal to the loop planes and periodic in the direction
of the conductor motion.
To reduce rotational losses an additional restriction was imposed on the field shape (Null-Flux
condition): the flux variation within each loop had to be zero in the equilibrium. This condition
lowered the average value of the electric field in the conductor, but the topology did not allow
making it zero everywhere. The satisfaction of the Null-Flux condition was extremely sensitive
to manufacturing inaccuracies.
This dissertation proposes a novel type of magnetic bearing stabilized by the field-conductor
interaction. In contrast to the other bearings based on this principle, the proposed design allows
exact satisfaction of the Null-E condition in the equilibrium regardless of the conductor shapes and
even in the presence of an axial loading. Because of this we refer to it as the Null-E Bearing. Null-E
Bearings also have potential for higher load capacity and stiffness than Null-Flux Bearings. Finally,
their performance is highly insensitive to the manufacturing inaccuracies.

ii
The Null-E Bearing in its basic form can be augmented with supplementary electronics to improve its performance. The power rating of the electronics in this case can be much smaller than
in conventional active magnetic bearings. Depending on the degree of the electronics involvement,
a variety of magnetic bearings can be developed ranging from a completely passive to an active
magnetic bearing of a novel type.
The dissertation contains theoretical analysis of the Null-E Bearing operation, including derivation of the stability conditions and estimation of some of the rotational losses. The validity of the
theoretical conclusions has been demonstrated by building and testing a prototype in which noncontact suspension of a 3.2-kg rotor is achieved at spin speeds above 18 Hz.

Acknowledgments
This work would not have been possible without the help of many people and many strokes of
good fortune. Therefore, in my acknowledgment section, I would like to look back on the way that
brought me to the idea of the magnetic bearing presented in this dissertation, and say thanks to God
and to all of the people who helped me in this long journey.
I first heard about magnetic bearings while studying navigational devices at Moscow N. E.
Bauman State Technical University (MSTU) in 1987. I consider myself fortunate that I have learned
the fundamentals of engineering from very talented professors teaching at MSTU. I am especially
grateful to Dr. Konovalov - an excellent engineer, manager, and a man of principles - who was a
real mentor for me.
The year I joined MSTU was also the year when High-Temperature Superconductivity was
discovered. Soon after, I joined a team working on the development of magnetic bearings for
navigational devices based on this effect. I greatly appreciate the help of Dr. Poluschenko, Dr.
Nizhelskii, Dr. Matveev and other members of this team who provided me with the support to
do my experiments on superconducting levitation. I am even more grateful to them for personal
support during difficult times.
In the following years, I worked on many other projects, which were more rewarding financially; however, I continued to work with Dr. Poluschenko and Dr. Nizhelskii on superconducting
magnetic bearings. I am thankful for their persistence and devotion to the science.
During my studies of superconducting bearings I understood much more about the physics of
magnetic levitation. This finally led me to a novel magnetic levitation principle, which does not rely
on superconductors, and which laid the foundation for this dissertation.
Because of the economical situation in Russia I could not continue my research, and in 1997 I
joined a British consulting company Ove ARUP and Partners. Here I met a very good friend of
mine, Adrian Salter, who became very interested in my principle of magnetic levitation. In 1999,
I left Ove ARUP and Partners. At that time Adrian and I filed an application for the US patent
on the magnetic bearing design. Adrian actively supported all my subsequent work on this project,
which would have been absolutely impossible without his help.
The same year I entered the University of Virginia, where I was very lucky to have Dr. Maslen as
my advisor. Not only did he allow me to pursue this idea as my Ph.D. project, but he also provided
intellectual and financial support. His broad expertise enabled him to understand my problems and
guide me in the right direction. Because of his continuous prodding I had to generalize the theory
of the bearing operation far beyond my original intent. This gave me many new insights.
The help of many other people was also essential for the success of my project. Dr. Gillies
helped me with my experiments and did a heroic work on proof-reading my papers. Lewis Steva
helped me with manufacturing the prototype. John Gray helped me with various practical issues.
I would also like to thank Carlos and Esther Farrar for awarding me their fellowship, and all my

iii

iv
advisory committee for their time, good words and valuable suggestions.
I am very grateful to everybody who gave me feedback on my work. In particular, I would like
to thank David Meeker for expressing his points of view on various aspects of my dissertation, good
insights, suggestions, and healthy criticism.
Further, I need to thank my friends John Gray, Michael Baloh and many others, for moral
support, strong drinks and good parties.
Finally, I must to say Thank You to my mom for raising me, and then supporting and encouraging me all the time. I would like to dedicate this work to her.

Alexei Filatov

Contents
1 Introduction
1.1 Problem definition and research motivation . . . . . . . . . . . . . . . . . . . . .
1.2 Summary of the dissertation . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3 Dissertation outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1
1
3
6

2 Literature review
2.1 Active Magnetic Bearings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Magnetic Bearings using tuned LC circuits . . . . . . . . . . . . . . . . . . . . . .
2.3 Diamagnetic and Superconducting Bearings . . . . . . . . . . . . . . . . . . . . .
2.4 Bearings utilizing conductor/magnet interaction . . . . . . . . . . . . . . . . . . .
2.4.1 Eddy-current bearings using AC field . . . . . . . . . . . . . . . . . . . .
2.4.2 Eddy-current bearings involving relative motion . . . . . . . . . . . . . . .
2.4.3 Null-flux bearings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5 Bearings using a combination of magnet/magnet andconductor/magnet interactions
2.6 Bearings stabilized by a gyroscopic torque . . . . . . . . . . . . . . . . . . . . . .
2.7 Position of the proposed electromagnetic bearing in the general classification . . .
2.8 Other relevant developments . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.9 Prior publications of the material presented in the dissertation . . . . . . . . . . . .

8
8
8
9
9
9
10
10
12
12
12
17
17

3 Theoretical Analysis
3.1 Conductor rotating about a fixed axis in an axisymmetric magnetic field
3.2 Dynamics of a rotating conductor with slow lateral motions . . . . . . .
3.3 Dynamics of a rotating conductor with damped lateral motion . . . . .
3.4 An inverse system with a stationary conductor and a rotating magnet. . .
3.5 5-DOF non-contact suspension . . . . . . . . . . . . . . . . . . . . . .
3.6 Effects of the stationary conductor inductance on system stability . . . .
3.6.1 Equations of motion . . . . . . . . . . . . . . . . . . . . . . .
3.6.2 Stability conditions . . . . . . . . . . . . . . . . . . . . . . . .
3.6.3 Analysis of the stability conditions . . . . . . . . . . . . . . . .
3.7 Rotational loss due to the rotor unbalance . . . . . . . . . . . . . . . .
3.7.1 Losses neglecting the damper inductance . . . . . . . . . . . .
3.7.2 Losses including the damper inductance . . . . . . . . . . . . .

18
18
29
34
35
39
40
41
45
46
53
54
57

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

CONTENTS

vi

4 Technical and Practical Considerations


4.1 Angular stability of the suspension . . . . . . . . . . . . . . . .
4.2 Designing the rotating conductors and stationary magnetic field .
4.3 Generation of the stationary magnetic field . . . . . . . . . . . .
4.4 Designing the stationary conductors and rotating magnetic field .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

59
59
60
72
73

5 Experiments
5.1 Observation of stable non-contact levitation . . . . . . . . . . . . .
5.1.1 Experimental arrangement . . . . . . . . . . . . . . . . . .
5.1.2 Results and discussion . . . . . . . . . . . . . . . . . . . .
5.2 Measurement of the radial suspension stiffness . . . . . . . . . . .
5.3 Measurement of the rotational losses . . . . . . . . . . . . . . . . .
5.4 Effects of the stationary coil inductance on the suspension properties
5.4.1 Influence of damper inductance on liftoff speed . . . . . .
5.4.2 Experimental arrangement . . . . . . . . . . . . . . . . . .
5.4.3 Results and discussion . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

78
78
78
81
81
83
89
89
90
91

.
.
.
.

6 Conclusions

95

A Exact Stability Bounds

105

B Force Calculations
107
B.1 Direct force calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
B.2 Force calculation using energy methods (Section 3.1) . . . . . . . . . . . . . . . . 111
C Magnetic Field Calculations
113
C.1 Magnetic system without pole shoes. . . . . . . . . . . . . . . . . . . . . . . . . . 114
C.2 Magnetic system with pole shoes. . . . . . . . . . . . . . . . . . . . . . . . . . . 116
C.3 Choosing parameters of the magnetic system. . . . . . . . . . . . . . . . . . . . . 119
D Loading Characteristics

122

E Experimental details
128
E.1 Measurement of the rotating coil inductances and resistances . . . . . . . . . . . . 128
E.2 Measurement of the radial position of the rotor. . . . . . . . . . . . . . . . . . . . 128
F Measured Loading Characteristics
G The high rotational speed assumption.
G.1 Avoiding the high spin rate assumption . . . . . . .
G.2 Bode plot analysis. . . . . . . . . . . . . . . . . .
G.2.1 Open loop model in the frequency domain.
G.2.2 Effects of on the system stability. . . . .

133

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

150
151
155
155
158

List of Figures
1.1
1.2

An example clarifying application of Earnshaws theorem . . . . . . . . . . . . . .


The structure of the proposed magnetic suspension. . . . . . . . . . . . . . . . . .

2.1
2.2
2.3
2.4
2.5
2.6
2.7

Explanation of the Null-E principle in eddy current bearings. . . . . . . . . .


The operational principle of a rotational null-flux suspension. . . . . . . . . .
The operational principle of an embodiment of the proposed suspension. . . .
Comparison to Null-Flux bearings: the rotor rotates about its symmetry axis. .
Comparison to Null-Flux bearings: the rotor rotates about an arbitrary axis. .
The inverse system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Further generalization of the scheme shown in Figure 2.3. . . . . . . . . . . .

.
.
.
.
.
.
.

10
13
13
14
14
15
16

3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8
3.9
3.10
3.11
3.12
3.13
3.14
3.15

Rotating conductor G exposed to a circumferentially uniform magnetic field Bz . .


Coordinate frames. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
An example clarifying why the force always passes through the axis Z0 . . . . . . .
Average electromagnetic force acting on body G. Rotation axis Z is fixed. . . . . .
Dependencies of the in-plane stiffness K and the angle on the rotational speed .
Orientations of the force components. . . . . . . . . . . . . . . . . . . . . . . . .
Diagram of the velocity dependent forces. . . . . . . . . . . . . . . . . . . . . . .
Inverse system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Explanation of the stabilizing effect due to the inductance of the stationary coils. .
Arrangement of stationary coils and a movable magnet. . . . . . . . . . . . . . . .
Possible mutual orientations of the graphs W2 (x) and W1 (x). . . . . . . . . . . .
Calculated stability regions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Analysis of the stability region . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Force diagram with rotor unbalance. Stationary coil inductances are neglected. . .
Force diagram with rotor unbalance. Stationary coil inductances included. . . . . .

18
19
23
27
28
33
35
36
41
42
50
54
54
56
57

4.1
4.2
4.3
4.4
4.5
4.6
4.7
4.8
4.9

An explanation of static angular stabilization . . . . . . . . . . . . . . . . . .


An explanation of stabilizing mechanism due to the gyroscopic effect . . . . .
A motivation for the choice of the conducting loop and magnetic field shapes. .
A particular case of the general structure shown in Figure 3.1. . . . . . . . . .
Cross-section of the loop and magnetic field as in Figure 4.3 in the XZ plane. .
XZ-plane cross-section with conducting slab instead of a loop . . . . . . . . .
Field distribution approximated by a third order polynomial. . . . . . . . . . .
Field distribution approximated by a third order polynomial plus point currents.
A conducting loop formed by two infinitely long round wires. . . . . . . . . .

59
60
61
61
62
62
65
66
67

vii

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

2
4

LIST OF FIGURES

viii

4.10
4.11
4.12
4.13
4.14
4.15
4.16
4.17

Geometric parameters of the conducting loops in Figure 4.4. . . . . . . . . .


Multi-turn loop with additional inductance. . . . . . . . . . . . . . . . . . .
Forming a multi-turn loop using PCB technology. . . . . . . . . . . . . . . .
Squirrel-cage rotor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A variant of the magnetic system . . . . . . . . . . . . . . . . . . . . . . . .
Magnetic field distribution in the system shown in Figure 4.14 . . . . . . . .
An electronic schematic to reduce the apparent resistance of a stationary coil.
Voltage distribution along the path of the current in the stationary coil. . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

69
71
71
72
73
73
74
74

5.1
5.2
5.3
5.4
5.5
5.6
5.7
5.8
5.9
5.10
5.11
5.12
5.13
5.14
5.15
5.16
5.17
5.18
5.19
5.20
5.21

Cross-sectional schematic view of the bearing prototype. . . . . . . . . . . . .


The stationary magnetic field distribution in the prototype. . . . . . . . . . . .
The rotor of the prototype. . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The rotor being suspended vertically. . . . . . . . . . . . . . . . . . . . . . . .
An overall view of the test rig. . . . . . . . . . . . . . . . . . . . . . . . . . .
Radial equilibrium force diagram. . . . . . . . . . . . . . . . . . . . . . . . .
Measured radial destabilizing stiffness Kdes introduced by the axial suspension.
Predicted and measured radial suspension stiffness, K. . . . . . . . . . . . . .
Predicted and measured angle . . . . . . . . . . . . . . . . . . . . . . . . . .
Rotor speed decay during run-downs in vacuum. . . . . . . . . . . . . . . . . .
Rotational power losses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Approximations of the measured rotational loss in the original prototype. . . .
Approximations of the measured rotational loss in the modified prototype. . . .
Stability regions for the bearing prototype . . . . . . . . . . . . . . . . . . . .
Comparison of the exact and approximate stability bounds . . . . . . . . . . .
Two variants of the connection of the winding of the additional inductor. . . . .
Measured coil voltage at the stability transition. . . . . . . . . . . . . . . . . .
Measured coil voltage when unstable. . . . . . . . . . . . . . . . . . . . . . .
Oscillograms measured during stable levitation. . . . . . . . . . . . . . . . . .
Oscillograms measured when the bearing operated in a limit cycle. . . . . . . .
Coil voltage spectra with and without additional inductance. . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

79
79
79
80
81
82
83
84
85
86
86
87
88
90
91
91
92
92
93
94
94

B.1 The operational principle of an embodiment of the proposed suspension. . . . . . . 107


B.2 Direct calculation of the forces acting on a conducting loop. . . . . . . . . . . . . 108
C.1
C.2
C.3
C.4
C.5
C.6
C.7

A variant of the magnetic system . . . . . . . . . . . . . . . . . . . . . . . . . . .


A model for the field calculation in the system without pole shoes. . . . . . . . . .
= 1/3 and g = 0.4.
Mid-gap field distribution assuming constant charge density.
A model for the field calculation in the system with pole shoes. . . . . . . . . . . .
Periodic expansion of the system with pole shoes. . . . . . . . . . . . . . . . . . .
= 1/3 and g = 0.4. . . .
Midgap field distribution assuming constant potential.
Comparison of the calculated and measured field distributions. . . . . . . . . . . .

113
114
116
116
117
120
120

D.1
D.2
D.3
D.4
D.5

Predictions of the stiffness per radial footprint area. . . . .


Predictions of the load capacity per radial footprint area. .
Predictions of the stiffness vs. bearing volume ratios. . . .
Predictions of the load capacity vs. bearing volume ratios.
Predictions of the stiffness vs. bearing weight ratios. . . .

123
124
125
125
126

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

LIST OF FIGURES

ix

D.6 Predictions of the load capacity vs. bearing weight ratios. . . . . . . . . . . . . . . 126
D.7 Predictions of the stiffness vs. rotor weight ratios. . . . . . . . . . . . . . . . . . . 127
D.8 Predictions of the load capacity vs. rotor weight ratios. . . . . . . . . . . . . . . . 127
E.1
E.2
E.3
E.4
E.5

AC voltage and current measured in a coil and an additional inductor. . . .


Sensor output voltage vs. distance dependence measured in the x-direction.
Sensor output voltage vs. distance dependence measured in the y-direction.
Sensor output voltage vs. distance dependence in the x-direction . . . . . .
Sensor output voltage vs. distance dependence in the y-direction . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

129
130
131
131
132

F.1
F.2
F.3
F.4
F.5
F.6
F.7
F.8
F.9
F.10
F.11
F.12
F.13
F.14
F.15
F.16
F.17
F.18
F.19
F.20
F.21

Crosscoupled force-displacement characteristic at 1320 RPM. . . . . . . . . . .


Direct force-displacement characteristic in the X-direction at 1600 RPM. . . . . .
Crosscoupled Force-displacement characteristic
in the Y-direction at 1600 RPM.

Force vs total rotor displacement, R = X 2 + Y 2 , characteristic at 1600 RPM. .


The angle between displacement and force vs rotor displacements at 1600 RPM.
Direct force-displacement characteristic in the X-direction at 1800 RPM. . . . . .
Crosscoupled force-displacement characteristic
in the Y-direction at 1800 RPM.

Force vs total rotor displacement, R = X 2 + Y 2 , characteristic at 1800 RPM. .


Angle between displacement and force vs rotor displacement at 1800 RPM. . . .
Direct force-displacement characteristic in the X-direction at 2000 RPM. . . . . .
Crosscoupled force-displacement characteristic
at 2000 RPM. . . . . . . . . . .

Force vs total rotor displacement, R = X 2 + Y 2 , characteristic at 2000 RPM. .


Angle between displacement and force vs rotor displacement at 2000 RPM. . . .
Direct force-displacement characteristic at 2200 RPM. . . . . . . . . . . . . . .
Crosscoupled force-displacement characteristic
at 2200 RPM. . . . . . . . . . .

Force vs total rotor displacement, R = X 2 + Y 2 , characteristic at 2200 RPM. .


Angle between displacement and force vs displacement at 2200 RPM. . . . . . .
Direct force-displacement characteristic at 2400 RPM. . . . . . . . . . . . . . .
Crosscoupled force-displacement characteristic
at 2400 RPM. . . . . . . . . . .

2
Force vs total rotor displacement, R = X + Y 2 , characteristic at 2400 RPM. .
Angle between displacement and force vs displacement at 2400 RPM. . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

134
136
136
137
137
139
139
140
140
142
142
143
143
145
145
146
146
148
148
149
149

List of Tables
3.1
3.2

Summary of the stability conditions. . . . . . . . . . . . . . . . . . . . . . . . . .


Important axes of the system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

53
55

5.1

Parameters used to predict the prototypes radial loading characteristics. . . . . . .

84

C.1 Suggested geometry of the radial suspension magnetic system with pole shoes. . . 121
D.1 Parameters of the rotating coils in the bearing prototype. . . . . . . . . . . . . . . 123
D.2 Parameters of the stationary magnetic system used in the radial suspension. . . . . 123
D.3 Densities of the major materials used in the radial bearing. . . . . . . . . . . . . . 124
E.1 Measured parameters of the rotating coils used in the bearing prototype. . . . . . . 129
F.1
F.2
F.3
F.4
F.5
F.6
F.7
F.8
F.9
F.10
F.11
F.12

Raw data for calculating the radial force-displacement characteristic at 1320 RPM.
Summary of the radial suspension parameters measured at 1320 RPM. . . . . . . .
Raw data for calculating the radial force-displacement characteristic at 1600 RPM.
Summary of the radial suspension parameters measured at 1600 RPM. . . . . . . .
Raw data for calculating the radial force-displacement characteristic at 1800 RPM.
Summary of the radial suspension parameters measured at 1800 RPM. . . . . . . .
Raw data for calculating the radial force-displacement characteristic at 2000 RPM.
Summary of the radial suspension parameters measured at 2000 RPM. . . . . . . .
Raw data for calculating the radial force-displacement characteristic at 2200 RPM.
Summary of the radial suspension parameters measured at 2200 RPM. . . . . . . .
Raw data for calculating the radial force-displacement characteristic at 2400 RPM.
Summary of the radial suspension parameters measured at 2400 RPM. . . . . . . .

133
134
135
136
138
139
141
142
144
145
147
148

Nomenclature
A
A
B
C
c
D
E
E
F
fL
f
G
G
H
I
i
J
Jz
j
K
k
L
l
M
M
m
N
n
P
P
Q
Q
R
S
s
T

Amplitude
Work
Magnetic flux density
Damping coefficient
Mass normalized damping coefficient
Determinant of a quadratic equation
Electrical field strength
Electromotive force
Mechanical force
Volume density of Lorenz forces
Reciprocal of the inductive coil time constant (units of frequency)
Designation of a rotating conductor
Designation of a stationary conductor
Magnetic field strength
Electrical
current

= 1
Current density per unit length
Polar moment of inertia about axis Z
Current density per unit surface area
Stiffness
Mass normalized stiffness
Inductance (or inductance matrix)
Length
Mutual inductance (or mutual inductance matrix)
Magnetization
Mass
Number of wire turns in a coil
Circular periodicity of an object
Pressure
Power
Designation of a magnetic field source mounted on the stator
Designation of a magnetic field source mounted on the rotor
Resistance (or resistance matrix)
Surface
area
= c/ k (dimensionless)
Time period

xi

NOMENCLATURE
T
U
u
v
Wm
Z
Z
Z0
Z0
Zar
Zas
Zm

r0
u

0
r

p
x

x
t

x
x

<>

Mechanical torque
Voltage
= k 2 (dimensionless)
Velocity
Electromagnetic energy
Rotor rotation axis
Stationary conductor symmetry axis
Stator magnetic field symmetry axis
Rotor magnetic field symmetry axis
Axial suspension axis on the rotor
Axial suspension axis on the stator
Axis on the rotor passing through its center of mass
Polar angle of a displacement vector
Polar angle of a velocity vector
Displacement of the Z axis from the axis Z0
Distance between the rotor center of mass center and the axis Z0 (unbalance)
Gain of a transfer function
Phase of a transfer function
Magnetic permeability
Magnetic permeability of vacuum
Relative magnetic permeability
Current vs displacement ratio (or a matrix composed of such ratios)
Root of a characteristic equation
Magnetic flux
Angle complimentary to the angle between a force and a displacement
Inductive coil time constant
Conductivity
Circular rotational frequency
Circular frequency as a variable in the frequency domain analysis
Rotor whirl speed
Rotor precession speed
Complex conjugate of x
Time derivative taken in the rotor coordinate frame
x is a parameter of the radial suspension affected by the axial suspension
x is a normalized length
Time averaging

xii

Chapter 1

Introduction
1.1

Problem definition and research motivation

Using electromagnetic forces to support rotating objects (rotors) without mechanical contact is an
appealing technical solution in many situations since it allows rotating systems to operate at much
higher speeds than obtainable using conventional mechanical bearings, and without risk of overheating and wear of the suspension components. Such suspensions must meet specific engineering
performance requirements including (typically): adequate load capacity and stiffness in given size,
low levels of rotational loss and external energy consumption. Sufficient load capacity and stiffness
can often be obtained simply using the interaction between permanent magnets and/or between permanent magnets and permeable materials. Thus, if two permanent magnets having magnetization
M are brought in contact, the magnetic pressure P on their surfaces is [1]
P =

(0 M)2
.
20

(1.1)

Recently developed relatively inexpensive and very powerful rare-earth magnets such as NdFeB [2]
have 0 M close to 1.2 T essentially independent of demagnetizing fields in a wide field range.
According to Eq. (1.1), this results in a magnetic pressure of 0.57 MPa. This value is high enough
to satisfy load capacity requirements for many applications. (For comparison, the typical maximum
operating pressures for fluid film bearings is about 2 MPa and for rolling element bearings it is
about 6 MPa). High values of stiffness can also be obtained through proper design of the magnet
arrangement [3]. Importantly, no external energy is required to produce a force using permanent
magnets, as long as no net work is done. In Figure 1.1, the repulsive force between two magnets is
used to compensate for the weight of the top magnet. Note that, if the magnetic field generated by
the top magnet is uniform circumferentially about its axis Z, there will be no drag torque exerted
about this axis.
Unfortunately, the physical limitation imposed by Earnshaws theorem [4] rules out the possibility of realization of stable static non-contact suspension using only permanent magnets and
permeable materials. Strictly speaking, this theorem was originally formulated for constant charges
(which could be electric, magnetic or gravitational charges) interacting with fields (electric, magnetic or gravitational), which could be described through a potential function obeying Laplaces
equation. In this case, the conclusion of the theorem is a direct consequence of the maximum principle for Laplaces equation [5] maintaining that the maximum or minimum of the potential function

CHAPTER 1. INTRODUCTION

Figure 1.1: This magnet arrangement could be used to compensate the weight of the top magnet,
but it is unstable radially.
always occurs on the boundary and never in free space. This theorem was later expanded by Braunbek [6, 7] to include materials with magnetic permeability and electrical permittivity greater than 1.
Both results are usually collectively referred to as Earnshaws theorem. This theorem can be neatly
formulated in terms of suspension stiffnesses K in three dimensions:
Kx + Ky + Kz 0

(1.2)

or, in linear algebraic terms, [8]: the trace of the lateral suspension stiffness matrix is zero or less
than zero. Recognizing that a necessary condition of the suspension stability is that each of the
diagonal elements of this matrix is positive, this implies that stability is not possible. Note that
Earnshaws theorem does not impose any restrictions on angular stiffnesses. In fact, suspension
stability about all angular degrees of freedom can be achieved in a passive magnetic system by
proper design (see a simple argument in Section 4.1). The main concern is, therefore, the lateral
stability of the levitated body.
For the axially symmetric suspension shown in Figure 1.1, Eq. (1.2) simplifies to
Kax + 2Krad 0

(1.3)

where Kax is the axial suspension stiffness and Krad is the radial suspension stiffness. From this
equation, it is apparent that if stable suspension is achieved in the axial direction (Kax > 0), it will
be unstable in the radial direction (Krad < 0) and vice versa. Thus the top magnet in Figure 1.1
will tend to slip to the side from the central equilibrium position.
There are, however, several well-known methods of achieving non-contact suspension using
electromagnetic forces in spite of the restrictions imposed by Earnshaws theorem. To summarize,
such methods make use of either diamagnetic or superconducting materials, conducting objects
interacting with time-varying magnetic fields, gyroscopic torques, or feedback control systems.
In rotational systems, only the last method has found application so far, because in contrast to
the other it can provide a high level of load capacity and stiffness, a low level of rotational losses and
a wide range of operating temperatures. The suspension systems utilizing this method are referred
to as Active Magnetic Bearings (AMB). However, they also have significant drawbacks including
requirement of rather complicated control systems, external power sources and connecting wires.
This often complicates integration of a bearing into a final device, for example if it has to operate
in a closed vessel or in vacuum. More discussion of the electromagnetic suspension methods is
provided below in the Literature Review.
In the present dissertation we investigate the potential of magnetic bearings utilizing the interaction of conducting objects with time-varying magnetic fields, or, more specifically, interaction

CHAPTER 1. INTRODUCTION

of conducting objects with non-homogeneous magnetic fields in relative motion. The known advantages of this type of magnetic bearing include simplicity, reliability, a wide range of operating
temperatures and system autonomy (no external wiring and power supplies are required). The disadvantages of the earlier embodiments were high rotational loss, low stiffness and load capacity.
It was realized, however, that rotational loss, load capacity and stiffness depend strongly on the
topology of the conductors and the magnetic fields.
In theory, the rotational loss in the equilibrium position in the absence of external loading can
be made zero by designing a system so that no electrical field E is induced in the conductor during
the rotor rotation. This is referred to as the Null-E condition in the dissertation. In the earlier
embodiments, the Null-E condition could not be satisfied exactly unless infinitely thin conductors
were used.
Load capacity and stiffness can be also maximized through choosing shapes of the conductors
and the fields. From this point of view the field and the conductor shapes used in the so called NullFlux Bearings were found to be advantageous: the conductors were shaped as planar loops with
central openings and the fields were orthogonal to the loop planes and periodical in the direction of
the conductor motion.
To reduce rotational losses at equilibrium, an additional restriction was imposed on the field
shape (Null-Flux condition): the flux variation within each loop had to be zero. This condition
lowered the average value of the electrical field in the conductor, but the topology did not allow
making it zero everywhere. The satisfaction of the Null-Flux condition was extremely sensitive
to manufacturing inaccuracies.
After analysis of the existing designs, we propose a novel topology of the conductor/field system. This topology takes advantage of the technical solutions leading to high load capacities and
stiffnesses in Null-Flux Bearings, but at the same time allows exact satisfaction of the Null-E condition regardless of the shapes of the conductors used. In fact the Null-E condition is satisfied not
only in the absence of external loading but also when a purely axial loading is applied. Because of
this, the bearing potentially has very low rotational losses, which would be essentially zero if the
loading is axial - a condition which is often easy to satisfy in stationary applications.
Considering all the above we named the proposed bearing a Null-E Bearing by analogy with
Null-Flux Bearings. Other important advantages of the Null-E Bearings include potentially higher
load capacity and stiffness than can be attained by Null-Flux Bearings and robustness to manufacturing inaccuracies.
The Null-E Bearing in its basic form can be augmented with supplementary electronics to improve its performance. The power rating of the electronics in this case can be much smaller than
that in Active Magnetic Bearings. Depending on the degree of the electronics involvement, a variety of magnetic bearings can be developed ranging from a completely passive to an active magnetic
bearing of a novel type.

1.2

Summary of the dissertation

The subjects of this dissertation are a novel method of non-contact electromagnetic suspension and
a novel class of magnetic bearings.
The proposed class of magnetic bearing features a combination of properties which are not
collectively provided by any of the prior technologies. These include:
1. Very high reliability due to intrinsic stability;

CHAPTER 1. INTRODUCTION

Figure 1.2: The structure of the proposed magnetic suspension.


2. Wide temperature operation range including room temperature;
3. High efficiency;
4. Load capacity and stiffness sufficient for many applications;
5. Low rotational loss, virtually zero if only axial loading is applied - a condition which is easy
to satisfy in stationary applications;
6. No connecting wires and power supplies, unless supplementary electronics are used to augment system performance;
7. If supplementary electronics are used, their power rating will be much lower than those in
active magnetic bearings;
8. Low price;
9. Low requirements for manufacturing accuracy.
10. Easy introduction of active components, permitting a continuum of design from fully passive
to fully active.
Further, the proposed suspension principle allows easy integration with external control systems
resulting in variety of magnetic bearing designs featuring different degrees of the active component
involvement.
The method is based on the combination of the static interaction between permanent magnets
and the dynamic interaction between conductors and permanent magnets. The former is used to
achieve axial suspension, while the latter is used to achieve radial stabilization. The moving conductors interact with an axially directed magnetic field which is also required to be circumferentially
uniform. The latter requirement results in rotational losses being virtually zero under any purely
axial loading provided that it does not exceed the maximum axial load capacity of the system.
Moreover, together with some other measures, it minimizes rotational losses under radial loading.
These properties exhibit very little sensitivity to variations of many system parameters including
dimensioning inaccuracies. The suspension is stable for any rotational speed above a certain critical
value.
Its main idea is schematically explained in Figure 1.2, showing the arrangement of two magnets
Q and Q as in Figure 1.1, but further including two conducting disks G and G attached to the
magnet surfaces. Static interaction between magnets Q and Q provides axial suspension, while
dynamic interaction between conducting disk G and magnet Q together with dynamic interaction

CHAPTER 1. INTRODUCTION

between conducting disk G and magnet Q stabilizes the system in the radial direction and provides
positive radial stiffness. Note, that both conductors G and G are necessary; the suspension is
unstable if only one of the conductors is used. It is easy to see that in the equilibrium position when
the axis of the top magnet coincides with the axis of the bottom magnet, no magnetic field variation
occurs at any point of the conducting disks and no electrical field is induced. This is referred to
as the Null-E condition throughout the dissertation. Apparently because of the Null-E operation
no currents are induced and no energy dissipation takes place in the conducting volumes when the
bearing is at equilibrium.
The dissertation includes a general analysis of the interaction in this type of system, considering
arbitrary magnetic means Q and Q , generating circumferentially uniform magnetic fields, and arbitrary conductors G and G . For simplicity, it is assumed that the magnetic fields are directed axially
within the conductors. This implies that Ampere forces acting on the conductors are purely radial
and do not influence the operation of the axial suspension. The analysis includes the following
steps:
1. Analysis of the radial electromagnetic forces acting on an arbitrary conducting body G exposed to a magnetic field Bz (r) generated by sources Q. The field is circumferentially uniform about some axis Z0 and is directed along this axis. The conductor rotates about some
axis Z parallel to Z0 . The sources Q are stationary.
The equations for the forces acting on the body G are derived.
2. Analysis of the dynamics of slow radial motions of body G superimposed on high speed
rotation about axis Z.
It is shown that if the slow motion assumptions are satisfied, the equilibrium that occurs when
axis Z coincides with axis Z0 is always unstable.
3. Analysis of the inverse system, including magnetic sources Q rotating about some axis Z0
and interacting with a stationary conductor G . The field generated by the sources Q is
circumferentially uniform about axis Z0 .
It is rather surprising, but these two structurally identical systems behave completely differently. The source of this difference is clarified. The equations for the forces acting on Q are
obtained assuming that the inductive properties of the body G can be neglected.
4. Analysis of the radial motion of a combined body Q + G comprising bodies Q and G
arranged so that axes Z0 and Z coincide, interacting with a combined body Q+G comprising
bodies Q and G arranged so that their axes Z0 and Z  coincide. This analysis does not
consider the axial suspension, which can be provided using the static interaction between the
magnetic field sources Q and Q (as in Figure 1.2), or between separate sets of magnets. It is
assumed that the center of mass of the body Q + G lies on the Z axis.
This time it is shown that there is an in-plane equilibrium position for which axis Z (Z0 )
coincides with axis Z  (Z0 ) and this equilibrium can be made stable. The stability condition
is derived.
5. An analysis that includes the axial suspension.
Here, a stability condition for the whole assembly is derived.

CHAPTER 1. INTRODUCTION

6. Modified stability analysis including inductive component of the impedance of the conductor
G . The analysis is carried out for a practically important case when the conductor G is
formed by a set of individual current loops with negligible mutual inductance.
It is shown that the inductive component may have a stabilizing effect and that the minimal stable levitation speed can be reduced through appropriate choice of the stationary coil
inductance.
While Figure 1.2 indicates the major components of our design, to fully realize its potential, the
shapes of the conductors and the fields need to be optimized. In particular, the ideas which were used
earlier to increase load capacity, stiffness and minimize energy dissipation in the out-of-equilibrium
operation in Null-Flux and Superconducting Bearings were found to be useful here. These and other
design considerations as well as some practical modifications to the device structure are discussed
farther in the dissertation. One of the significant modifications is using external circuitry to control
currents in the stationary conductors G .
The dissertation also includes an estimation of the rotational losses caused by the non-coincidence
of the rotor center of mass and the Z0 axis equivalent of unbalance in this bearing. This loss model
does not include the effects caused by the axial suspension.
A significant part of the dissertation is devoted to the analysis of systems involving either conductors rotating in stationary circumferentially uniform magnetic fields, or sources of such fields
rotating about the field symmetry axis in the presence of stationary conductors. This analysis has a
value of its own, since systems of this type can be encountered in other electromechanical devices
such as electromechanical dampers.
To verify theoretical results, a suspension system has been built and tested, in which a 3.2-kg
rotor is suspended without mechanical contact when it rotates above approximately 18 Hz. The
dissertation describes the actual hardware and the tests conducted to establish its conformance with
the theory.

1.3

Dissertation outline

Chapter 1 gives the motivation and objectives of the dissertation research followed by a brief dissertation summary.
Chapter 2 is a comprehensive survey of literature on various types of magnetic bearings along
with discussion of advantages and disadvantages of each particular type. The chapter indicates the position of the principle developed in this dissertation within the context of general
classifications of magnetic suspension technologies.
Chapter 3 contains a theoretical analysis of the electromagnetic interactions encountered in the
proposed system, system dynamics and stability.
Chapter 4 suggests some practical solutions which could be useful in designing magnetic bearings
based on the proposed principle.
Chapter 5 describes the experimental part of the work aimed to verify the theoretical results.
Chapter 6 concludes the dissertation with a summary of the obtained results and suggestions for
future research.

CHAPTER 1. INTRODUCTION

Appendix A includes the MathCAD script used to calculate exact stability bounds taking into account the inductance of the stationary coils.
Appendix B includes an example of using the general theory developed in this dissertation to calculate parameters of a particular embodiment of the Null-E Bearing, in which the conductor
G is represented by a set of conducting coils.

Chapter 2

Literature review
There are several well-known methods of achieving non-contact suspension using electromagnetic
forces in spite of the restrictions imposed by Earnshaws theorem. Each method has its advantages
and disadvantages, which are discussed in this chapter.

2.1

Active Magnetic Bearings

Active Magnetic Bearings (AMB) employ control systems to vary the supporting magnetic field
in response to rotor displacements in order to produce restoring forces. For this field to be easily
variable, it has to be generated by current-carrying coils [9,10,11,12,13,14,15,16,17,18,19,20,21,
22, 23, 24, 25, 26]. Active Magnetic Bearings can produce high load capacity and stiffness, but they
also have significant drawbacks, some of which originate from the very principle of their operation.
Indeed, compensation of high loads requires strong magnetic fields associated with large amounts
of energy. If the loads are dynamic, these fields have to be varied quickly. This requires powerful
electronics. Partially because of the high power rating, AMB controllers are relatively expensive,
bulky and can be unreliable. Another practical disadvantage of AMBs is that they require external
power sources and connecting wires. This often complicates their integration into a final device, for
example if the bearing has to operate in a closed vessel or in vacuum.

2.2

Magnetic Bearings using tuned LC circuits

This type of magnetic bearing ( [27], [28]) is, in fact, very close to AMBs in the sense that it also
relies on coils with currents which vary depending on the rotor position. Consequently, it also
requires external wiring and power supplies. The advantages, as compared to most AMBs, are that
LC bearings are simple and do not have position sensors. The disadvantages are principally much
lower efficiency, load capacity and stiffness. Furthermore, in contrast to an AMB, the parameters of
this type of bearing cannot be tuned easily. Finally, Active Magnetic Bearings can be also designed
so as to operate without an explicit position sensor (these are so called Self-Sensing Bearings,
see, for example, [16]). Because of these reasons, LC bearings can hardly compete with AMBs in
most cases.

CHAPTER 2. LITERATURE REVIEW

2.3

Diamagnetic and Superconducting Bearings

Another approach to achieve stable non-contact suspension is to utilize magnetic substances to


which Earnshaws theorem is not applicable: diamagnetic materials or superconductors. The superconductors in some cases can also be viewed as perfectly diamagnetic materials almost completely
expelling magnetic fields from their interior [29, 30]. This phenomenon is known as the Meissner
effect and can be observed in Type I superconductors and Type II superconductors below the first
critical field value. Type II superconductors above the first critical field value allow the field to enter,
but if they belong to the category of so called Hard Type II superconductors, further variations of
the field are suppressed.
Forces acting on ponderable masses of any known diamagnetic materials such as graphite or
bismuth are negligible unless magnetic fields on the order of tens of teslas are used [31, 32, 33, 34,
35, 36] which cannot be obtained with permanent magnets. This makes such bearings impractical
for most commercial applications.
The interaction of a superconductor with a permanent magnet may result in significant forces
[37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55, 56, 57, 58]. However, the requirement of cooling superconductors to cryogenic temperatures prevents the use of this type of bearing
in most applications. Moreover, if the Hard Type II superconductors are used and if they interact
with magnetic fields exceeding the first critical value, the bearing force-displacement characteristics exhibit strong force-displacement hysteresis. The latter makes the rotor equilibrium position
unpredictable and may cause bearing failure under vibrations [52]. At the same time, use of Hard
Type II superconductors working above the first critical field value results in much higher load capacity and stiffness than obtainable with Type I superconductors. This is because Hard Type II
superconductors can carry much higher currents in much stronger magnetic fields. Moreover, recently discovered High-Temperature superconductors [59, 60] such as YBaCuO (which are typical
Hard Type II superconductors) remain superconducting at much higher temperatures (well above
the boiling point of liquid nitrogen) than known Type I Superconductors. Because of the above,
significant efforts have been made to design magnetic bearings using Hard Type II and, particularly, High-Temperature superconductors. Thus, a method was found to reduce force-displacement
hysteresis significantly [55, 56, 57, 58]. The same method allowed obtaining relatively high load capacity and stiffness per unit surface area of the bearing due to optimal usage of the superconductor
current-carrying capacity and the energy of the permanent magnets. Some results of this work are
applied in this dissertation to the design of a bearing which does not use superconductors.

2.4
2.4.1

Bearings utilizing conductor/magnet interaction


Eddy-current bearings using AC field

Similar to diamagnetic and superconducting materials, normal conductors such as copper suppress
AC magnetic fields within their volume. Thus a non-contact suspension can be obtained using
the interaction of a conductor with an AC magnetic field. This method has found application in
containerless melting, where a conductor suspended in AC fields melts due to heating by eddy
currents [61,62,63,64,65,66,67,68,69,70]. The advantage of non-contact suspension in this case is
that the crucible material does not contaminate the melt. While, in this particular application, high
losses are beneficial (they are needed to melt the conductor), in other applications the suspension,
in contrast, has to be highly efficient.

CHAPTER 2. LITERATURE REVIEW

10

Figure 2.1: Explanation of the Null-E principle in eddy current bearings.

2.4.2

Eddy-current bearings involving relative motion of a conductor and


nonhomogeneous field

Note that energy in the previous type of bearing dissipates in both the suspended conductor and the
coils generating the supporting AC field. The latter are eliminated in another embodiment, in which
the supporting non-homogeneous magnetic field is generated by permanent magnets. This field is
time-invariant in the magnet coordinate frame, but seen as time-varying by a conductor moving with
respect to it. The early designs of such bearings exhibited low stiffness and load capacity and high
rotational loss [71, 72].
It was realized, however, that the rotational loss in the no-load equilibrium can be made essentially zero by designing a system so that no magnetic field variation occurs in the conductor during
the rotor rotation in this position [73]. Consequently, no electrical fields and no currents will be
developed. Figure 2.1 clarifies this point using a linear bearing example. It is easy to see that the
magnetic field on the middle line of this structure is zero and when the conductor is thin, the field is
close to zero within the conductor volume. Motion of the conductor along the middle line, therefore,
does not cause significant electrical fields, currents and losses.
In this dissertation we refer to the requirement of zero electrical field within the conductor
volume during its motion in the no-load equilibrium position as the Null-E condition even though
this term, to our knowledge, is not used in the literature. Note that the Null-E condition cannot be
satisfied exactly within the conductor shown in Figure 2.1 unless the conductor is infinitely thin.
Even an approximate satisfaction of this condition requires high geometrical accuracy, especially
in rotational systems. In contrast, the bearing structure proposed in this dissertation allows exact
satisfaction of the Null-E condition regardless of the conductor shape.

2.4.3

Null-flux bearings

Other disadvantages of the early eddy-current bearings were low load capacity and stiffness. Again,
it was realized that this is not an inherent problem of the method: both load capacity and stiffness
could be increased dramatically through careful choosing the topologies of the conductors and the
fields. Low values of these parameters in the eddy-current bearings were caused by an inefficient usage of the conductor current-carrying capacities and the magnet energies. The eddy currents within
the conducting volumes were distributed and oriented with respect to the magnetic fields rather arbitrarily, resulting in arbitrarily oriented Lorenz forces. This caused unwanted resistive losses and
stresses in the conductors. In the ideal scenario, the current at each point of the conducting volume
should be orthogonal to the field and directed so that the resulting Lorenz force is directed oppositely to the disturbing force acting on the rotor. Besides, it is desirable to concentrate the magnetic
fields generated by the magnets in compact areas and force the currents to flow in these areas. All
these measures would lead to increases of the load capacity, stiffness and their ratios with respect to

CHAPTER 2. LITERATURE REVIEW

11

the energy dissipating in the conductor when it is displaced from the equilibrium. They were major
guidelines in developing so called Null-Flux Bearings.
The Null-flux principle was first proposed for application to linear suspension (high-speed
trains). In the late 1960s, Danby and Powell [74] suggested using a high-speed train suspension
scheme comprising an array of magnets of alternating polarity installed on a train and generating
flux trough a set of stationary, vertically oriented coils. The flux was required to vary monotonically
with the levitation height, passing through zero at some point. Consequently, this type of suspension
is generally referred to as Null-Flux suspension. It is easy to see that if the train moved constantly
at the height corresponding to the zero flux, there would be no macroscopic currents induced in the
coils. However, if the levitation height varied, currents would appear because the magnet polarity
alternates in the direction of motion. This would result in lifting and drag forces. This scheme has
been proven to deliver much higher lift/drag ratios ( 60/1) than the scheme in which simple coils
or aluminium slabs laid on the ground interact with a high-gradient magnetic field generated on the
train (lift/drag 25/1) [75].
The principle of null-flux suspension is implemented in Japanese MagLev trains [76], [77]. Significant theoretical work in this direction has been done in Argonne National Laboratory [78], [79]
and by Davey [75, 80, 81, 82]. The null-flux suspension concept is also being developed by FosterMiller (USA) with application to both high-speed trains [83] and space vehicle launch systems [84].
It is easy to see that the system topology in the Null-Flux Bearing does not allow satisfaction of
the Null-E condition in the equilibrium: the parts of the conducting loops orthogonal to the motion
will experience time-varying magnetic fields. The Null-Flux condition is weaker than the Null-E.
It is rather surprising, but the null-flux principle has been extended to rotational systems only
recently [85, 86, 87, 88]. One of the significant problems with direct application of this technology
to the rotational systems is that in this case the null-flux condition is especially difficult to satisfy
because it requires very high geometrical accuracy of the bearing components. (The author earlier
has successfully applied the methods of increasing suspension stiffness and load capacity developed
for the Null-Flux Bearings to the design of a superconducting rotational bearing [55, 56, 57, 58].
In this dissertation we point out that there is a way to extend the null-flux principle to rotational
systems other than doing it directly, as in [85,86,87,88]. As a result we obtained a novel non-contact
suspension principle, which has much in common with null-flux suspensions, but also has significant
differences. This principle is applicable only for rotational systems and cannot be adapted to linear
ones. It allows exact satisfaction of the Null-E condition (which is a stricter condition than the
Null-Flux one) and because of this we refer to it as a Null-E suspension. Importantly, the Null-E
condition is satisfied not only in the absence of any loadings, but also when the loading is purely
axial - a condition that is often easy to satisfy in stationary applications.
One of the other advantages of the bearing proposed in this dissertation is potentially higher
load capacity and stiffness than in Null-Flux Bearings. The reason for this is that in the Null-Flux
Bearings only the time-averaged Ampere force acting on each conducting coil is directed oppositely
to the rotor displacement, while there is a time period when the instantaneous force is directed in the
direction of the displacement [85]. Moreover, since the field in a Null-Flux Bearing is required to
be circumferentially periodical, its magnitude and, consequently, the Ampere force are not always
of the maximal magnitude.
In contrast, in the proposed bearing, the field is circumferentially uniform, and when the bearing
is exposed to a constant loading it responds with a time-invariant force opposite to the loading, as
will be shown later in the dissertation.

CHAPTER 2. LITERATURE REVIEW

2.5

12

Bearings using a combination of magnet/magnet and


conductor/magnet interactions

A passive magnetic bearing structure can be developed combining the static interaction between
permanent magnets and the dynamic interaction of a conductor with a magnetic field. The first type
of interaction provides compensation at constant loading, while the second stabilizes an otherwise
unstable system [89, 73]. If a structure like that shown in Figure 2.1 is used as a stabilizer, then the
overall system can be designed so that there would be no energy dissipation not only in the absence
of any loading, but also when the loading is purely axial [73]. For example, this loading can be the
rotor weight if the bearing axis is vertical. Unfortunately, this feature is very sensitive to bearing
parameter variations, especially dimensioning inaccuracies. Also, the design of the subsystem using
a conductor interacting with a magnetic field as suggested in [89, 73] is essentially an eddy current
bearing with all the subsequent disadvantages: high rotational loss when the rotor is displaced from
the equilibrium and low stiffness and load capacity.
One of the major contributions of this work is the development of an electromagnetic stabilizer
which offers significant advantages compared to both Eddy-Current and conventional Null-Flux
Bearings, including lower rotational losses and higher stiffness and load capacity.

2.6

Bearings stabilized by a gyroscopic torque

Note that the suspensions utilizing interaction of conductors with permanent magnets work only in
the dynamic mode, i.e., when the rotor rotates. It was shown recently, that a dynamically stable
suspension can be realized using only permanent magnets [90, 91, 92, 93, 94, 95]. This does not
contradict Earnshaws theorem which is valid only for stationary systems and does not consider
dynamic effects caused by rotation. Unfortunately, this method was found difficult to use in practice
because of low stiffness and load capacity and because the stability is very sensitive to variations of
many parameters including rotor weight, residual magnetization of the magnets and, importantly,
rotation speed: the suspension is stable only within a narrow speed range limited from both below
and above.

2.7

Position of the proposed electromagnetic bearing in the general


classification

The approach most similar to the one proposed in this dissertation is the null-flux suspension. In
principle, the proposed scheme can be viewed as a special zero-order case of a rotational null-flux
suspension, however, there are significant differences between the two schemes which should be
kept in mind. The following discussion briefly examines the major similarities and differences.
The operational principle of a rotational null-flux suspension is indicated in Figure 2.2. When
the rotor axis of symmetry is in the central position, the flux through the conducting coil is required
to be zero. It is easy to see that when this axis is displaced from the central position, the flux through
the coil can be described in terms of a Fourier cosine series expansion, as follows:
= 01 cos(t)


k=1

Ak cos(kn0 t + k )

(2.1)

CHAPTER 2. LITERATURE REVIEW

13

Figure 2.2: The operational principle of a rotational null-flux suspension.

Figure 2.3: The operational principle of an embodiment of the proposed suspension.


where n0 is half of the number of magnet blocks around the circumference (n0 = 2 in Figure 2.2);
Ak and k are the amplitude and phase of the harmonic with frequency kn0 . The parts of the
bearing are interchangeable: the magnets can be mounted on the rotor while coils can be made
stationary.
An embodiment of a part of the suspension proposed here can be obtained from the suspension
Figure 2.2 if only one magnet of each polarity is used as shown in Figure 2.3. (We specify that this
is a part of the suspension here because by itself it does not define an equilibrium position of the
rotor as discussed below.) The flux through the coil in this case can be represented as
= 02 cos(t)

(2.2)

This equation can be obtained from (2.1) by taking n0 = 0.


While the schemes shown in Figure 2.2 and Figure 2.3 look very similar, there is a significant
difference between them. In Figure 2.2, there is a unique point of equilibrium, when the rotor symmetry axis coincides with the stator symmetry axis. In contrast, in the system shown in Figure 2.3,
regardless of what is the axis about which the rotor spins, there will be no currents and no forces
induced as long as this rotation axis coincides with the field symmetry axis.
Figure 2.4 and Figure 2.5 clarify this difference. Figure 2.4 shows a rotor with four conducting
loops exposed to a circumferentially uniform magnetic field. This structure is a generalization of
Figure 2.3. The density of the dashed circles represents the magnetic flux density: the field has

CHAPTER 2. LITERATURE REVIEW

14

Figure 2.4: Rotor axis of symmetry = Rotation axis = Field axis of symmetry: no currents and no
forces.

Figure 2.5: Rotor axis of symmetry = Rotation axis = Field axis of symmetry: again no currents
and no forces.

CHAPTER 2. LITERATURE REVIEW

15

Figure 2.6: The inverse system: a source of the field is mounted on the rotor, a conductor is stationary.
to be radially non-uniform. It is easy to see that when the rotor symmetry axis coincides with the
magnetic field symmetry axis and the rotor rotates about this axis there will be no currents and,
subsequently, no forces induced. The same would be observed in the null-flux bearing (Figure 2.2)
if the geometry of the coils is chosen to meet the null-flux requirement.
Assume, however, that in the system shown in Figure 2.4 the rotor rotates as a whole about
another axis, which is not its axis of symmetry. This can be absolutely any axis which, in particular,
can be located outside of the rotor as shown in Figure 2.5. Note, that in this case there still will be no
changes of the magnetic fluxes through the coils and, correspondingly, no currents and forces. The
system shown in Figure 2.3 and Figure 2.5 does not define an equilibrium as a particular position
of the rotor with respect to the stator. Instead, it defines a kind of a pseudoequilibrium position
where the rotor rotation axis coincides with the magnetic field symmetry axis. The rotor rotation
axis is defined arbitrarily. Apparently, small perturbations of the rotor motions accumulated over
time will redefine the rotor rotation axis: the system shown in Figures 2.3 and 2.5 by itself does not
prevent the rotor from moving around.
In contrast, it is easy to see that if in the system shown in (Figure 2.2) the rotor rotated about
some axis different from its symmetry axis, then forces would develop tending to align the rotor
symmetry axis (moving in circles) with the stator symmetry axis.
The situation with the system shown in Figure 2.3 and 2.5 becomes even worse when rotor
inertia is taken into account. In this case, there is a unique equilibrium when the rotor center of
mass coincides with the rotation axis which, in its turn, coincides with the magnetic field symmetry
axis. However, as one can anticipate and as is shown in this dissertation, this equilibrium is unstable
and the rotor center of mass would tend to spiral away from it. Therefore, the system shown in
Figures 2.3 and 2.5 is not functional on its own.
It becomes functional, however, when augmented with a second subsystem, which is structurally
identical to the first but has the source of the circumferentially uniform magnetic field mounted on
the rotor so that the field symmetry axis is located in proximity of the rotor center of mass, while the
conductor is mounted stationary. We refer to this system as the inverse one in order to distinguish
it from the direct system including a conductor rotating in a stationary field. An example of the
inverse system is shown schematically in Figure 2.6.
It is easy to see that rotation of the magnet about its symmetry axis Z does not cause any currents
in the stationary coils regardless of the position of the magnet. At the same time lateral motions of
the magnet do cause currents and, consequently, forces. If the inductive properties of the stationary

CHAPTER 2. LITERATURE REVIEW

16

Figure 2.7: Further generalization of the scheme shown in Figure 2.3.


coils are neglected, then the force is dissipative in nature and the system acts as a selective damper
suppressing all the motions of the rotor but the rotation about the Z axis.
Summing up, the system shown in Figures 2.3 and 2.5 dictates that in equilibrium the rotation
axis of the rotor coincides with the symmetry axis of the stationary magnetic field, but does not
determine where this axis passes through the rotor. The system shown in Figure 2.6 dictates that
the rotation of the rotor would take place about the symmetry axis of the magnet mounted on the
rotor but does not tell how this axis has to be located with respect to the stator. Only both systems
together define a unique equilibrium. If the symmetry axis of the magnet shown in Fig. 2.6 passes
through the rotor center of mass, then the point of equilibrium is defined as being where this axis
coincides with the stationary magnetic field symmetry axis. Otherwise, the dynamic balance of
the centrifugal and damping forces acting on the rotor will define (uniquely) some other axis of
the rotor which has to coincide with the stationary magnetic field symmetry axis for equilibrium to
occur. Thus, the proposed system requires magnets and conductors mounted on both the rotor and
the stator to define the equilibrium position of the rotor with respect to the stator.
In contrast, conventional Null-Flux Bearings need only one part of the suspension to define the
equilibrium, which includes either conductors mounted on the rotor and magnets mounted on the
stator or, oppositely, the magnets mounted on the rotor and conductors mounted on the stator.
(It is interesting to note that the rotational null-flux suspension (Figure 2.2) when unrolled produces a linear suspension. It is easy to see, however, that unrolled suspension shown in Figure 2.3
is not functional: there is no force acting on a coil moving along a magnetic field that is uniform in
the direction of motion.)
Note that when the proposed bearing is in equilibrium, no change of the magnetic field occurs
in either rotating or stationary conductors and, consequently, no electrical field is induced (NullE condition is satisfied) regardless of the shapes of the conductors. The only requirement is that
the magnetic fields have to be uniform circumferentially. Thus Figure 2.7 shows an arbitrarily
shaped conductor G instead of the conducting loop shown in Figure 2.3 and Figure 2.4. This a huge
practical advantage since it minimizes the requirements for manufacturing accuracy.
The major components of the bearing in assembly are indicated in Figure 1.2 with the stationary
and rotating conductors represented by conducting discs attached to the magnet surfaces.
As shown in the dissertation, the proposed magnetic bearing has several important practical
advantages compared to Eddy-Current and conventional Null-Flux Bearings.

CHAPTER 2. LITERATURE REVIEW

2.8

17

Other relevant developments

Rotation of a conductor in a circumferentially uniform magnetic field, which is the basis of the
suspension principle described here, has already been studied to some extent with application to
electromagnetic eddy-current dampers. Such a damper typically includes either a conducting disc
mounted on the rotor and interacting with a stationary circumferentially uniform magnetic field or,
inversely, a source of such a field mounted on the rotor and interacting with a stationary conductor
[96, 97, 98, 99, 100, 101].
Several authors [97, 98] have indicated that electromagnetic eddy-current dampers are very effective under certain operating conditions. However, conditions (generally at high speed) in which
a rotating system becomes unstable have also been documented [99, 100].
It was noticed that system behavior at high speeds is very different depending on whether the
conductor rotates while the field source is stationary, or inversely, the field source rotates while the
conductor is stationary [100]. One of the explanations [100] was based on the well known general
fact that the effect of damping on the whirl motion of a rotor is always stabilizing if the energy
dissipates in the stationary part of the damper; however, if the energy dissipates in the rotor, the
effect is stabilizing for all whirl motions at a frequency above the rotational speed and destabilizing
in other cases [102].
A detailed explanation of this phenomenon has been given in [101], where it was shown that the
destabilizing effect in the case of a rotating conductor occurs due to cross-coupling stiffness terms.
The analysis was carried out for a thin, electrically conductive, but non-ferromagnetic disk attached
to the rotor and exposed to a circumferentially uniform magnetic field normal to the disc surface. It
was assumed that the disc magnetic Reynolds number is low: Rm << 1.
If the conductor G in Figure 2.7 were disk-shaped as in [101] then, in contrast, the operation
of the proposed bearing would require Rm 1. The analysis of this case is carried out in the
dissertation (another parameter designated as angle was introduced instead of Rm to characterize
an arbitrary conductor G). It is shown that, besides the destabilizing cross-stiffness, a stabilizing
direct radial stiffness also develops but, nevertheless, the system remains unstable. It becomes
stable, however, if augmented with the inverse system including a field source mounted on the rotor
interacting with a stationary conductor. Further, it will be shown here that under certain conditions,
the inverse system also provides additional stabilizing radial stiffness and cannot be viewed simply
as a damper.

2.9

Prior publications of the material presented in the dissertation

A part of the material presented in the dissertation can be also found in several journal and conference papers published during the work on this project: [103, 104, 105, 106, 107]. The proposed
method of non-contact suspension and a magnetic bearing based on this method are protected by
US Patent No. 6,304,015 [107].

Chapter 3

Theoretical Analysis
3.1

Conductor rotating about a fixed axis in an axisymmetric


magnetic field

Consider a solid conductor G of arbitrary shape exposed to an external magnetic field. This magnetic field is required to be circumferentially uniform about some axis Z0 and non-uniform in the
radial direction. For simplicity, the flux density is also required to be parallel to the axis Z0 , i.e. it
can be expressed as Bext = Bz = Bz (r)ez , where ez is a unit vector directed along the axis Z0 .
Further, it is assumed that the conductor rotates about some axis Z parallel to Z0 with angular speed
(Figure 3.1).
Our first goal is to find the forces acting on the rotating conductor assuming that axis Z is fixed
and displaced from axis Z0 by some very small distance r0 = {x0 , y0 }. Clearly, if r0 = 0,
there are no currents and no forces due to the system rotational symmetry.
Consider the case when r0 = 0.
The following coordinate frames will be used (see Fig. 3.2):
1. A stationary coordinate frame X0 Y0 Z0 ;
2. A coordinate frame X1 Y1 Z moving laterally (but not rotating) together with the rotor;
3. A coordinate frame XR YR Z linked firmly with the rotor.
The equations of mechanical motion of body G will be analyzed in an inertial coordinate frame
X0 Y0 Z0 while equations for the electromagnetic field will be analyzed in a rotating coordinate

Figure 3.1: Rotating conductor G exposed to a circumferentially uniform magnetic field Bz .


18

CHAPTER 3. THEORETICAL ANALYSIS

19

Figure 3.2: Coordinate frames.

frame XR YR Z related to conductor G. This can be done conveniently since the influence of acceleration on electromagnetic processes is known to be negligible for most practical electromechanical
systems. [108].
Let us consider some point A within the conducting volume of the rotor, which location is given
by polar coordinates and in the rotor coordinate frame XR YR Z, and by and in the X1 Y1 Z
coordinate frame.
The distance from the point A to the axis Z0 is
r = {(x0 + cos )2 + (y0 + sin )2 }1/2

(3.1)

Since r0 is small, we replace (3.1) with a Taylors series expansion, accurate to the second order
terms:
r = + x0 cos + y0 sin

(3.2)

Change of r:
r = r = x0 cos + y0 sin

(3.3)

The magnetic flux density at point A is


Bz (, , r0 ) = Bz () +

dBz
r
dr

or, using (3.3)


Bz (, , r0 ) = Bz () +

dBz
(x0 cos + y0 sin )
dr

(3.4)

If the rotor rotates with an angular speed , then = + t, and Bz at each point of the
conductor G varies in time.
Differentiating (3.4) with respect to time, we get
z
dBz
B
=
(x0 sin t + y0 cos t)
t
dr
where indicates that the derivative is taken in the rotor coordinate frame.

(3.5)

CHAPTER 3. THEORETICAL ANALYSIS

20

Using polar coordinate representation of the displacement vector


x0 = r0 cos ;

y0 = r0 sin

we rewrite (3.5) as
z
dBz
B
=
r0 sin(t + (t))
t
dr

(3.6)

The distributions of current density, j, and the strength of the magnetic field produced by this
current, Hj , within the conductor G are given by the following quasi-stationary approximations of
Maxwells equations:
Hj = j

E =

(3.7)
z
Hj
B

t
t

(Hj ) = 0

(3.8)
(3.9)

where E is the electric field strength, = r 0 , 0 is the magnetic permeability of vacuum and r
is the relative magnetic permeability of a given material.
In addition we need to consider a constitutive equation linking j and E (Ohms law):
j = E

(3.10)

in which is the electrical conductivity. In the most general case, the properties and in equations
(3.8), (3.9) and (3.10) can be functions of spatial coordinates.
z
B
in Eq. (3.7) through Eq. (3.7) are resolved in the
Note that all the vectors j, Hj , E and t
rotor coordinate frame XR YR Z.
Combining (3.7), (3.8) and (3.10) we get an expression for the strength of the magnetic field
induced by the current in conductor G (field diffusion equation):


Hj
1


( Hj ) +
= B
(3.11)
z (r, r0 , t)ez

t
t
This equation together with (3.9), subject to the boundary conditions, gives us Hj . Knowing
Hj , we can further find j using equation (3.7).
We can be more specific about the boundary conditions here. Since equation (3.11) allows for
arbitrary spatial distributions of and , this equation, in principle, can be used to describe field
and current distributions in any rotor. (This would not be the most practical way to attack the
problem in most cases, however. In practice, the rotor can be divided in several areas with constant
and , so that the solutions can be obtained for each area separately, and then coupled together
through boundary conditions. The analysis is even easier if the rotor consists of several insulated
bulk conducting coils and iron cores.) Since the bearing is supposed to be non-contact, the solution
inside the rotor should be coupled to the solution in free-space through boundary conditions on the
rotor surface: normal component of the current, j, is zero, tangential component of Hj and normal
component of Bj are continuous. The magnetic field in free space can be described through a scalar

CHAPTER 3. THEORETICAL ANALYSIS

21

potential given by Laplaces equation subject to the boundary conditions at infinity (the field should
be zero there).
The only information about the boundary conditions which we will really need for further analysis, however, is that they are linear and time-invariant. Then, the operators acting on Hj on the
left-hand sides of equations (3.9) and (3.11) are linear and time-invariant. Further, the current density j can be obtained from H j using a linear transformation given by (3.7). We summarize all
these observations by writing in a general form
L,,z,t (j) =

z
B
ez
t

(3.12)

where L,,z,t is a linear operator over variables , , z, t.


For our analysis, we represent two-dimensional vectors by complex numbers with real parts
corresponding to the x-components and imaginary parts corresponding to y-components. We mark
the complex representations of the vectors with hats:
0 = x0 + iy0 = r0 ei
r
Using this form, we rewrite (3.6) as
z
dBz   i(t+) i/2 
B
=
Re r0 e
e
t
dr
Considering that the real part of a complex vector a can be represented as Re(a) =
where * means complex conjugation, we rewrite (3.13) as

(3.13)
1
2 (a

+ a ),

z
1 dBz   i(t+) i/2  i(t+) i/2 
B
=
r0 e
e
+ r0 e
e
t
2 dr

(3.14)

We know that when a linear system described by (3.12) is subjected to a harmonic excitation
z
B
= Aei0 eit
t

(3.15)

the resulting current density is also a harmonic function:



j = ei Aei0 eit

(3.16)

where A is a constant amplitude of the input signal, 0 is a constant signal phase, is an amplification factor and is a phase shift. Thus, the current density in the conductor is
z
B

j = ei
t
Knowing the current density, one can easily find the volume density of the Lorenz force:
fL = j Bz
Because Bz is directed along the Z axis, the fL vector is in the radial plane and it is normal to
the current density vector j. This can be easily reflected using complex vector representations by
multiplying the current density by ei/2 :
i/2

f
L = jBz e

CHAPTER 3. THEORETICAL ANALYSIS

22

or

i/2 i Bz
f
e
L = Bz e
t
1
i/2 i dBz 
0 ei(t+) ei/2 }
{r0 ei(t+) ei/2 + r
e
(3.17)
f
L = Bz e
2
dr
The equation (3.17) gives the projections of Lorenz force on the axes of the rotating coordinate
frame XR YR Z. To obtain the projections on the axis of the stationary coordinate frame X0 Y0 Z0
(represented by a complex number f
L0 ), the complex representation of the force in the rotating

frame fL (3.17) has to be multiplied by eit :
it
f
L0 = fL e
Thus,
1
i/2 i dBz 
0 e2it ei ei/2 }
{r0 ei ei/2 + r
f
e
L0 = Bz e
2
dr
or
1
i dBz 
0 ei e2it }
{r0 ei r
f
L0 = Bz e
2
dr

(3.18)

The total force acting on the conductor can be found by integration of (3.18) over the conductor
volume:


(3.19)
f
F0 =
L0 dv
G

We can make a few interesting observations about this system, which we shall call lemmas.
Lemma 1:
The electromagnetic force acting on body G is always directed along a line passing
through the field symmetry axis Z0 .
In other words, for any current distribution in conductor G, there is never a torque about
axis Z0 . Also, considering that we can move a vector of a force along its line of action,
we can say that the force acting on body G is always applied at some point on axis Z0 .
Proof:
If there is some current with density j flowing in body G, a torque T acting on body G
about the axis Z0 due to this current can be found as

Rmax

r j Bdv = ez
Bz (r)r
j n ds dr,
T =
V

S(r)

where S(r) is a cylindrical surface concentric with axis Z0 , n is the unit vector normal
to the surface S(r), and Rmax is a radius of a cylindrical surface concentric with axis
Z
0 and surrounding all of the conductor G. The first integral is taken over body G and
j n ds is the current flow through a cylindrical surface S(r). Because j = 0 ,
S(r)
S(r) j n ds = 0, and therefore T = 0.

CHAPTER 3. THEORETICAL ANALYSIS

23

Figure 3.3: An example clarifying why the force always passes through the axis Z0 .
The proof of this Lemma can be clarified by a simple example. First, assume that the conducting
body G is just a conducting loop of arbitrary shape carrying current I (see Figure 3.3). Consider
an annulus with inner radius r and a differential thickness dr concentric with the Z0 axis. We refer
to the inner surface of this annulus as S(r). It is easy to see that the differential tangential forces
acting on the conducting parts located within the annulus are |F 1 | = |F 2 | = IBz (r)dr and they
produce opposite torques about the axis Z0 . Thus, the net torque is zero, and this is because the
field is uniform circumferentially and also because the current which crosses the surface S(r) in the
outward direction is exactly equal to the return current crossing this surface in the inward direction.
This is, in fact, true for any current distribution, as implied by the integral form of j = 0.
Lemma 2:
The time averaged electromagnetic force Fav is proportional in magnitude to the
rotor displacement with some proportionality coefficient K and directed at some
angle 0 /2 with respect to a vector opposite to the displacement vector
r0 . The angle opposes the rotation direction.
In particular this implies that the projection of this force on the displacement direction
is either negative or zero.
Proof:
Assume that a constant external magnetic field Bz is generated by a system Q of m current loops each carrying constant current IQi , where i = 1 . . . m. Such a current loop
system can also represent permanent magnets with residual magnetization independent
of magnetic field strength [1].
We can further represent conducting body G as a set of n current-carrying loops. Both
n and m can tend to infinity.
Let LQ and LG be inductance matrices for the current loop sets Q and G, which are

CHAPTER 3. THEORETICAL ANALYSIS

24

known to have the following structure

L<S>11 M<S>12
...
M<S>1r

..
M<S>21
...
...
.
L<S> =

..

.
...
...
M<S>(r1)r
M<S>r1
...
M<S>r(r1)
L<S>rr

where < S > can be either Q (in which case r = m) or G (in which case r = n),
L<S>ii represent self-inductances of conducting loops and M<S>ij represent mutual
inductances between the loops within the system of interest. Matrices L<S> are known
to be symmetric and positive-definite [109].
For the whole system including both bodies Q and G, a similar inductance matrix can
be presented as


LG MGQ
L=
,
MQG LQ
where MGQ and MQG are n m and m n matrices of mutual inductances between
T .
systems Q and G and MGQ = MQG
For one revolution of body G around axis Z, from conservation of energy we have the
following equation:

T
Text dt
Pres dt = 0
(3.20)
Ae + Wm +
0

where Ae is the work of external electromotive forces in system Q against electromotive forces induced by variations of external fluxes, Wm is the change of electromagnetic energy, Pres is the power of resistive losses due to the currents flowing in body G,
Text is the external torque acting on body G about axis Z (in the case of accelerating
rotation, Text will include an inertia term Jz ,
where Jz is the moment of inertia of
body G about axis Z) and T is the period of rotation of body G. We did not include
resistive losses in system Q, which remain constant independent of what is happening in the system because the currents in system Q stay constant. These losses must be
compensated by some extra work done by external sources, which are also not included
in (3.20).
Wm can be written as [109]:
1 T
1 T
T
LQ IQ + IG
LG IG + IQ
MQG IG
Wm = IQ
2
2

(3.21)

In order to the keep the currents in system Q constant, external sources have to do an
amount of work against the induced electromotive forces E equal to

T
T
T dQ
T
T
dt = IQ
IQ
Edt =
IQ
Q = IQ
(MQG IG ) (3.22)
Ae =
dt
0
0
where
Q = MQG IG

(3.23)

CHAPTER 3. THEORETICAL ANALYSIS

25

represents the magnetic fluxes induced in the loops of system Q by currents flowing in
system G.
As stated earlier, the components of the current density j are harmonic functions with
frequency equal to the rotational frequency of body G. In this discrete model, this
means that the components of IG are harmonic functions with frequency . Considering this, and the fact that both Wm and Ae are single-valued functions of IG , we can
conclude that, after a complete rotation, they both vanish. Therefore, from (3.20) we
have

T
Text dt =
Pres dt
(3.24)
0

That is, during one complete revolution, the resistive losses in body G are compensated
by the work of any external torques.
Now, using the principle of virtual work, we can find the projection of the average
electromagnetic force acting on body G on the displacement direction Fav r as
Fav r =

< Wm + Ae >
r

(3.25)

where Wm is the change of the magnetic field energy caused by the virtual displacement r (variation of r0 ), Ae is the work of the external sources that keep the currents in system Q constant, and the brackets <> mean time averaging.
Considering that IQ is kept constant, the variation of the magnetic energy (3.21) will
be
T
T
Wm = IG
LG IG + IQ
(MQG IG )

(3.26)

and the work against the electromotive forces caused by r is


T
T
Q = IQ
(MQG IG )
Ae = IQ

(3.27)

Substituting (3.26) and (3.27) into (3.25) we get


Fav r =

T L I >
< IG
IG
G G
T
= < IG
LG
>.
r
r0

(3.28)

We know that the components of the current density j are harmonic functions with amplitudes proportional to r0 . This implies that the components of IG are harmonic
functions with amplitudes also proportional to r0 . Let us introduce vector G , composed of the proportionality coefficients between components of IG and r0 such that
IG = G r0

(3.29)

Then (3.28) can be rewritten as


Fav r = < TG LG G > r0

(3.30)

TG LG G is a quadratic form of G with a positive-definite weighting matrix LG ,


which is positive at any instant of time. Therefore, whenever we displace the axis Z

CHAPTER 3. THEORETICAL ANALYSIS

26

from axis Z0 , the projection of the average force acting on body G in the direction of
displacement r0 , Fav r , as given by equation (3.30), will always be directed towards
axis Z0 (oppositely to r0 ).
From equation (3.24), assuming that the variation of rotational speed during one
revolution is negligible, we can find the average torque applied by the external forces
to be

T
1
1
Pres dt =
Pres dt 0.
(3.31)
< Text >=
T 0
2 0
Here PRes is
T
RG IG
Pres = IG

(3.32)

which is a quadratic form of IG with positive definite resistance weighting matrix RG .


Using equation (3.29),
Pres = TG RG G r02
Then (3.31) becomes
< Text

r02
>=
2

(3.33)

TG RG G dt.

(3.34)

The direction of this torque coincides with the direction of rotation, . External forces
do positive work to compensate for the resistive losses in the conductor G (Eq. (3.24)).
From Newtons second law we conclude that the average torque produced by internal
electromagnetic forces opposes the rotation . Earlier, we have shown that the electromagnetic force acting on body G is always applied at some point on axis Z0 and it
always has a component directed opposite to the displacement direction. In order to
produce an average drag torque about axis Z given by (3.31) or (3.34), the averaged
force also has to have a tangential component magnitude of which is

< Text >


r0 T T
r0
< TG RG G >
=
G RG G dt =
(3.35)
Fav =
r0
2 0

and which is directed so as to produce a torque about the axis Z opposing the rotation.
In other words, the average electromagnetic force has to be deflected from the direction
opposing the displacement direction by some angle 0 < /2 so that the angle
opposes the rotation direction as shown in Figure 3.4. The magnitude of the angle is
= arctan

Fav
1 < TG RG G >
= arctan
=
Fav r
< TG LG G >

 T
1 0 TG RG G dt
= arctan

T T LG G dt
0

(3.36)

The proportionality coefficient K between the magnitudes of the average force Fav and
the rotor displacement r0 is


2
1
T
T
< G RG G >
K = < G LG G >2 +
(3.37)

Thus, we have proven this Lemma.

CHAPTER 3. THEORETICAL ANALYSIS

27

Figure 3.4: Average electromagnetic force acting on body G. Rotation axis Z is fixed.
Note that the angle can be zero if and only if the power of the resistive losses Pres is zero. This
would be possible only if body G was made of a superconducting material. For non-superconducting
materials we have the following lemma.
Lemma 3:
K and are functions of the rotational speed . When increases, K() monotonically increases towards a saturation value, while () monotonically approaches
zero as arctan(1/).
Proof:
To see that K() monotonically increases with towards some saturation value, first
note that the excitation function on the right-hand side of the Eq. (3.11) is a harmonic
function with amplitude proportional to (Eq. (3.5)). The solution Hj of a linear
partial differential equation (3.11) also is a harmonic function with frequency . The
amplitude of its time derivative Hj /t is proportional to . When is low, this
term is small and the right-hand
side of
 equation (3.11) is balanced primarily by the
1
left-hand-side term ( Hj ) . Thus, at low , Hj increases with almost
linearly. With further increase of , the contribution from the other left-hand side term,
Hj /t becomes more and more significant, until finally this term dominates and
balances the right-hand side term, which is also proportional to . Because of this
mechanism, when increases, the amplitude of Hj and, correspondingly, the current
density j in body G monotonically increases towards a certain bound. Using Eq. (3.37),
it is easy to see that K() behaves in the same way.
To see that () monotonically approaches zero when increases, consider Eq. (3.36).
When varies, the vector G scales proportionally, and the ratio between the nominator and the denominator in the right-hand part of (3.36) does not change. Therefore,
when increases monotonically decreases approaching zero as arctan(1/).
The lemma is proven.

The dependencies of K and on are indicated schematically in Fig. 3.5.


Now, consider a particular structure of a conducting body G, which features periodicity with
respect to rotation about axis Z with period 2/n. We will refer to this structure as having rotational
periodicity of order n. An order of periodicity can be infinite if body G is circumferentially uniform
about axis Z (for example, a conducting disk concentric with axis Z). For the case, where body G
has a periodic structure we can prove the following lemma.

CHAPTER 3. THEORETICAL ANALYSIS

28

Figure 3.5: Dependencies of the in-plane stiffness K and the angle on the rotational speed .
Lemma 4:
If a conducting body G has an order of periodicity about axis Z which is at least
3, the electromagnetic force acting on body G for an infinitesimally small r0 will
be time invariant.
Strictly speaking this statement may be true for any r0 (not necessarily small) if the
geometry of the system is such that currents induced in conductor G remain harmonic
functions of time.
Proof:
A body G, which has an n-th order periodicity about axis Z, can be viewed as consisting of n identical bodies Gk located uniformly around the rotation axis Z. In this case
we can limit integration in (3.19) to one of the bodies Gk only. The forces acting on the
other bodies would differ only by a phase shift. The overall force acting on the whole
body G can be thus found as
0 =
F

Gk



n

1
i dBz
i 
i
2i[t+ 2
(k1)]


n
Bz e
ne r0 + r0 e
dv
e
2
dr
k=1

Using the equation for the sum of geometric series,


n


e2i[t+ n (k1)] = e2it

k=1
4

n 3,

ei n 1 = 0

and therefore,
n

k=1

e2i[t+ n (k1)] = 0

e4i 1
4

ei n 1

CHAPTER 3. THEORETICAL ANALYSIS


0 =
F

Gk

29

1
dBz
0 dv
Bz ei()
n r
2
dr

Thus, the force acting on the rotating conductor G in this case is indeed independent of
time.
Designate
i

Ke

= n
Gk

1
dBz
Bz ei()
dv
2
dr

(3.38)

Then
r = Kei r
0
F

(3.39)

This equation shows that the force acting on the rotating conductor G, which has rotational periodicity of at least 3-rd order, is time invariant and its magnitude is proportional to the displacement r0 of the rotational axis Z from the magnetic field symmetry axis Z0 with some proportionality coefficient K. Also, the vector of this force
is deflected from the vector opposing the displacement vector r0 by some angle .
From Lemma 2, 0 /2 and the angle opposes the rotation direction.
The lemma is proven.

3.2

Dynamics of a rotating conductor with slow lateral motions

In the previous section, we made several conclusions about the force acting on conducting body G
when axis Z is fixed in space. In this section, we analyze the in-plane motions of body G and, in
particular the stability of the equilibrium when axis Z coincides with axis Z0 . We will assume that
body G has rotational periodicity of at least third order about the rotation axis Z and that this axis
passes through the mass center of body G.
This last assumption implies zero centrifugal force, which would be a time varying force with
circular frequency equal to , and both assumptions together imply a constant restoring force acting
on body G whenever axis Z is displaced from axis Z0 by a constant distance. Notice that, if body G
is suspended freely, it tends naturally to rotate about its mass center. In practice, body G may have
rotational periodicity about some other axis displaced from the rotation axis Z passing through the
mass center. This is likely to happen due to manufacturing inaccuracies. Even in this case we may
note that there still will be an equilibrium when axis Z coincides with axis Z0 and there will be no
currents and no forces acting at this equilibrium. However, when axis Z is displaced from axis Z0 ,
the system would respond with a force having some time varying component.
For simplicity, we assume here that the body G has rotational periodicity exactly about axis Z
passing through the mass center.
First, we derive the expressions for the forces acting on the conductor G when its rotation
axis Z moves, i.e. r0 = r0 (t). Assume that the axis Z moves with a non-zero velocity
v0 = r0 = (x 0 , y 0 ). Then, the equation (3.5) becomes
z
dBz
B
dBz
=
(x0 sin t + y0 cos t) +
(x0 cos t + y0 sin t)
t
dr
dr

(3.40)

CHAPTER 3. THEORETICAL ANALYSIS

30

In polar coordinates:
x0 = r0 cos ;

y0 = r0 sin ;

x 0 = v0 cos ;

y 0 = v0 sin

we rewrite (3.40) as
z
dBz
dBz
B
=
r0 (t) sin(t + (t)) +
v0 (t) cos(t + (t))
t
dr
dr

(3.41)

Using the complex representation of the velocity vector:


0 = x0 + iy0 = r0 ei ;
r

v0 = x0 + iy0 = v0 ei

we rewrite (3.41) as
z

dBz   i(t+) i/2
B
=
Re r0 (t) e
e
+ v0 (t) ei(t+)
t
dr

(3.42)

Considering that the real part of a complex vector a can be represented as Re(a) = 12 (a + a ),
where * means complex conjugation, we rewrite (3.42) as
z
1 dBz  
B
0 (t) ei(t+) ei/2 +
=
r0 (t)ei(t+) ei/2 + r
t
2 dr

i(t+)

+ v0 (t)e

(3.43)

i(t+)

+ v0 (t) e

In the previous section, we made use of the fact that when a linear system described by (3.12)
is subjected to a harmonic excitation
z
B
= Aei0 eit
t
the resulting current density is also a harmonic function:

j = ei Aei0 eit
where A is a constant amplitude of the input signal, 0 is a constant signal phase, is an amplification factor and is a phase shift (see Eqs. (3.15) and (3.16)) .
In this section, however, input signal amplitudes r0 , v0 and phases and (see (3.41)) are
not constants, but functions of t. It makes our analysis significantly more difficult. To proceed, we
make a simplifying assumption that even though A and 0 are functions of time, they vary slow
compared to eit . Then an approximate solution of (3.12) can be found as

j = ei A(t)ei0 (t) eit

(3.44)

Quantitatively, a condition of A(t) and 0 (t) varying slow compared to eit can be presented as
0 << and dA
dt << A.
With application to our problem, we need to assume that
r0 << r0 ;

v0 << v0 ;

<< ;

<<

(3.45)

CHAPTER 3. THEORETICAL ANALYSIS

31

Note, that r0 << r0 does not imply v0 (t) cos(t+(t)) << r0 (t) sin(t+(t))
and, therefore, we need to consider both terms in (3.41).


2
2
0 )2 , where a0 is the acceleration
0 ) and a0 = v0 2 + (v
Considering that v0 = r0 + (r
of the rotor axis Z, the above conditions can be replaced with slightly more restrictive
v0 << r0

(3.46)

a0 << v0

(3.47)

The conditions (3.45), (3.46) and (3.47) imply that variations of r0 , v0 , and in time (lateral
motions of the rotor) are slow compared to the rotational speed about the axis Z. We will discuss
later how the analysis based on the slow lateral motion assumption can be used in practice. For now
we assume that the conditions (3.46) and (3.47) are satisfied. Then, repeating the derivation from
the previous section, we find the current density in the conductor and the Lorentz force density to
be
z
B

;
j = ei
t

z
B
i/2

f
= Bz ei/2 ei
L = jBz e
t

Substituting (3.43) into the above equation for the Lorenz force density, we get

1
i/2 i dBz 
0 (t) ei(t+) ei/2 +
e
f
r0 (t)ei(t+) ei/2 + r
L = Bz e
2
dr

+ v0 (t)ei(t+) + v0 (t) ei(t+)

(3.48)

The equation (3.48) gives the projections of Lorenz force on the axes of the rotating coordinate
frame XR YR Z. To obtain the projections on the axis of the stationary coordinate frame X0 Y0 Z0
(represented by a complex number f
L0 ), the complex representation of the force in the rotating
(3.48)
has
to
be
multiplied
by
eit :
frame f
L
it
f
L0 = fL e
Thus,

1
i/2 i dBz 
0 (t) e2it ei ei/2 +
f
e
r0 (t)ei ei/2 + r
L0 = Bz e
2
dr

+ v0 (t)ei + v0 (t) e2it ei

or

1
i dBz
0 (t)ei + r
0 (t) e2it ei +
f
r
L0 = Bz e
2
dr
+ v0 (t)ei ei/2 + v0 (t) e2it ei ei/2

(3.49)

The total force acting on the conductor can be found by integration of (3.49) over the conductor
volume:


F0 =
f
L0 dv
G

CHAPTER 3. THEORETICAL ANALYSIS

32

Now, as in the previous section, we consider a particular case when the body G consists of n
identical bodies Gk located uniformly around the rotation axis Z and limit integration to one of
the bodies Gk only. The forces acting on the other bodies would differ only by a phase shift. The
overall force acting on the whole body G can be thus written as

(k1)]
 ei #n e2i[t+ 2
0 (t)ei + r

n
+
nr

0
k=1
1
i dBz

Bz e
dv
F0 =

dr
Gk 2

+nv (t)ei ei/2 + v ei/2 ei #n e2i[t+ 2


(k1)]
n
0
0
k=1
Using the equation for the sum of geometric series,
n


e2i[t+ n (k1)] = e2it

k=1

e4i 1
4

ei n 1

ei n 1 = 0

n 3,
and therefore,
n


e2i[t+ n (k1)] = 0

k=1

0 =
F

Gk

1
dBz
0 }dv
Bz ei()
{nv0 ei/2 n r
2
dr

We see that the force acting on the rotating conductor G consists of two components: positiondependent Fr ,

1
dBz
0 n
r = r
Bz ei()
dv
(3.50)
F
dr
Gk 2
and velocity-dependent Fv :

i/2

n
Fv = v0 e

Gk

1
dBz
Bz ei()
dv
2
dr

(3.51)

Using the designation introduced in (3.38)

1
dBz
GBz ei()
dv
Kei = n
dr
Gk 2
we present the forces acting on the rotating conductor in the following form:
0 ei
r = K r
F

(3.52)

v = K v0 ei ei/2
F
(3.53)

Note, that Fr is exactly the force which we obtained in the previous section assuming the axis
Z is fixed. We already know that this force is oriented at angle 0 < < /2 with respect to
the vector opposing the displacement r0 and the angle opposes the rotation vector. Using this
information and comparing the equations (3.53) and (3.52), it is easy to see how the vector of the

CHAPTER 3. THEORETICAL ANALYSIS

33

velocity-dependent force is oriented with respect to the velocity vector v0 . Verbally, it is oriented at
angle /2 with respect to the vector opposing the velocity vector v0 . The direction of this angle
coincides with the rotation direction. The orientations of the position-dependent force component
with respect to the displacement vector and the velocity-dependent force component with respect to
the velocity vector are shown in Fig. 3.6. Note that the displacement and velocity vectors themselves
can be arbitrarily oriented with respect to each other.

Figure 3.6: Orientations of the force components.


It is easy to see that in this system we may have a purely restoring force (when = 0), but no
damping, or a purely damping force (when = 2 ), but no restoring force. In either case the system
cannot be stable (at least asymptotically). A more evolved analysis shows that this system is always
unstable (even in sense of Lyapunov) if > 0. To show this, we introduce new designations

x1 = x

x2 = x
y =y

1
y2 = y
For convenience, we also define
k=

K
m

where m is the rotor mass. With these designations, we write the systems equations of motion:

x 1 = x2

x 2 = k cos x1 k sin x2 k sin y1 + k cos y2


(3.54)
y 1 = y2

y 2 = k sin x1 k cos x2 k cos y1 k sin y2


or in matrix form

x 2
= [A]
y

y 2

x1
x2
y1
y2

0
1
0
0
k cos k sin k sin k cos

.
A=

0
0
0
1
k
k
k sin cos k cos sin

(3.55)

Following the standard Routh-Hurwitz procedure [110], one can show that the system described
by the equation (3.55) is always unstable for any > 0.
Therefore, for now we conclude that the system is always unstable if the slow lateral motion
conditions (3.46) and (3.47) are satisfied.

CHAPTER 3. THEORETICAL ANALYSIS

3.3

34

Dynamics of a rotating conductor with damped lateral motion

As a next step, we assume that the lateral motions of the conductor G are externally damped, i.e.
there exists a damping force applied at the conductor mass center:
Fd = Cv0

(3.56)

or in complex form:
d = C v0
F

(3.57)

Adding this force to the equations of motion (3.55), we get

x 1

x 2
= [A]

y 1

y 2

x1
x2
y1
y2

0
1
0
0
k

k cos c k sin k sin

cos
.
A=

0
0
0
1
k
k
k sin
cos
k cos c sin
(3.58)

where c = C/m.
The characteristic equation of the matrix A (Eq. (3.58)) is
(
(
'k
'
) k *2
k
4 + 2
sin + c 3 + c2 +
+ 2 c sin + 2k cos 2 + 2ck cos + k 2 = 0

Note, that if
c >>

(3.59)

then this characteristic equation becomes


4 + 2c3 + (c2 + 2k cos )2 + 2ck cos + k 2 = 0,
which is the same equation we would obtain by simply neglecting the velocity dependent component (equation (3.53)). Therefore, in this case we can use a simplified system of equations:

x 1

x 2
= [A]
y 1

y 2

x1
x2
y1
y2

0
1
0
0
k cos c k sin 0
.
A=

0
0
0
1
k sin
0 k cos c

(3.60)

Using the Routh-Hurwitz procedure [110], it is easy to show that this system is stable if
c>

sin
k
cos

(3.61)

A geometrical explanation of the fact that the velocity-dependent component of the electromagnetic force (3.53) can be neglected is that this force introduces only a small deviation of the damping
force from the direction opposite to the velocity direction as can be seen in Figure 3.7.

CHAPTER 3. THEORETICAL ANALYSIS

35

Figure 3.7: Diagram of the velocity dependent forces.


When a strong external damping is present, the velocity dependent force component (equation (3.53)) also could be neglected even if the slow motion condition (3.47) was not satisfied. (This
must be done with a certain amount of caution, however, since in this dissertation we do not derive a
strict quantitative criteria for when this can be done ). Therefore, in the presence of strong external
damping the slow motion condition (3.47) may not be considered.
Nevertheless, the other condition (3.46) cannot be ignored. Note, that it clearly cannot be satisfied at r0 = 0. Assume that we can find a bound on the speed of the rotor lateral motions:
vmax (this bound exists since the system is stable if (3.61) is satisfied). Then the condition (3.46) is
satisfied if r0 > R = vmax /. Thus, we can use our model in the region r0 > R. The higher
, the smaller R. If the condition (3.61) is satisfied, the region r0 > R is a basin of attraction.
This is the kind of stability which will be implied further when we say that the system is stable.
Recognizing that K and are both functions of , (3.61) implies that, for a given damping
coefficient, there exists a value of above which the rotor radial equilibrium is stable. The bound
(3.61) goes to infinity when approaches /2 and goes to zero when approaches 0.
Note that the amount of damping in this system does not determine the boundary between
assymptotic stability and stability in the sense of Lyapunov as in a mass-spring system, but instead
determines the boundary between the stability in the above sense and the lack of any stability.
In summary, we have shown with some simplifications, that in the presence of external damping,
the equilibrium that arises when the rotation axis Z coincides with the magnetic field symmetry axis
Z0 can be stable.

3.4

An inverse system with a stationary conductor and a


rotating magnet.

In the previous section we showed that the equilibrium where the rotation axis of the conductor
G coincides with the field symmetry axis Z0 can be stable if the lateral motions of the conductor
are damped. For applications, such as flywheels, it is desirable that lateral damping does not cause
rotational drag. In this section we show that this can be realized if an inverse system is used, which
also includes a conductor G and a source of a circumferentially uniform magnetic field Q , but this
time the conductor is stationary, while the field source rotates about the field symmetry axis.
For example, in Figure 3.8, body Q is represented by a disk-shaped, axially-magnetized permanent magnet rotating about its symmetry axis Z0 . It is rather surprising, but rotation of body
Q about axis Z0 in this system leads to completely different results than rotation of body G about
axis Z in the first system. The following explanation may, however, clarify this apparent paradox.
Notice that in both cases there are two important axes: the axis of relative rotation of two bodies
and the axis of magnetic field symmetry, and we are analyzing slow relative lateral motions of the
bodies superimposed on high-speed relative rotation about the rotation axis. The difference is that,

CHAPTER 3. THEORETICAL ANALYSIS

36

Figure 3.8: Inverse system.


in the second case, the rotation axis always coincides with the field symmetry axis regardless of the
mutual positioning of bodies Q and G , while in the first case it does not.
As with the first system, we can make several observations about the system behavior which we
will call lemmas. In this section, we neglect the inductive properties of the conductor Q .
Lemma 1:
The electromagnetic force acting on body Q is always applied at some point on
axis Z0 .
In other words, there is no torque about axis Z0 .
Proof:
Identical to the proof of Lemma 1 with further implication of Newtons second law.
Lemma 2:
The absolute value of the electromagnetic force acting on body Q is proportional
to the velocity of its lateral motion. If the velocity is zero, there is no force acting
on body Q regardless of its position.
Proof:
It is easy to see that, in contrast to the first system, rotation of body Q about its symmetry axis Z0 does not cause any currents in conductor G because the magnetic field
is circumferentially uniform about axis Z0 and the rotation always takes place about
this axis. Correspondingly, the equation for the time derivative of the magnetic field at
some point A in body G does not involve a component caused by rotation anymore:
z
dBz
B
=
(r) r cos A
t
dr

(3.62)

where r is the velocity of body G with respect to body Q (it is equal but opposed to the
velocity of body Q ), A is the angle between the vector drawn to point A from axis Z
and vector r.
Thus, in this system, currents in the conductor G and, correspondingly,
the force acting on it are proportional to r.
According to Newtons third law, the force
acting on body Q has the same magnitude but opposite direction.
Lemma 3:

CHAPTER 3. THEORETICAL ANALYSIS

37

The electromagnetic force acting on body Q always has a component opposing r.

Proof:
Existence of the electromagnetic force implies the presence of currents in conductor G
and therefore energy dissipation because of the non-zero resistance of the conductor.
According to conservation of energy, the resistive losses in conductor G in a unit time,
Pres , have to be equal to the work of external forces Fext acting on the moving body
Q in the same unit time. (In the case of accelerating motion, the external forces would
include the inertial forces: m
r .) Thus, Fext r = Pres > 0. Considering that,
according to Newtons third law, the electromagnetic force balances the external force,
this inequality implies the statement of this lemma.
Lemma 4:
If a conducting body G has periodicity of at least third order about some axis Z  ,
then whenever axis Z  coincides with axis Z0 , the electromagnetic force acting on
body Q is directed strictly oppositely to the velocity of lateral motion of body Q .
Proof:
Because of the periodicity assumption, conductor G can be thought as consisting of
n bodies Gk (k = 1 . . . n) located uniformly about axis Z  , which can be electrically
connected or insulated from each other.
It has been established that whenever any current flows in a conductor exposed to a
circumferentially uniform magnetic field, the resulting force passes through the field
symmetry axis (see Lemma 1). For simplicity, we assume that the field is directed
axially. Because of this, the Ampere forces are located in the radial plane. We introduce
a Cartesian coordinate system X  Y  Z  with the X  axis directed along the line of action
of the force produced by one of the conducting bodies Gk . Designating this body
as 1 and numbering the other bodies counterclockwise, we note that the force vector
produced by the k-th body is directed at the angle 2(k 1)/n with respect to the X 
axis. Thus the complex representation of the force vector produced by the k-th body is
 = F  ei2(k1)/n
F
k
k

(3.63)

We will also need local coordinates to describe a position of a point of interest A within
an individual conducting body Gk . As such we will use polar coordinates rA and A
with the origin of the local coordinate system being the origin of the global X  Y  Z 
coordinate system. For the body G1 , we will measure the angle A from the axis X  .
For an arbitrary body Gk , we will measure this angle from an axis rotated by the angle
2

n (k 1) counterclockwise from the axis X (force action line).
Assume that the magnet moves laterally with some velocity v directed at an angle
with respect to the X  axis. The complex representation of this vector is v = vei .
At some point A in a conducting body G1 , we have
z
dBz
B
=
(rA )v cos(A )
t
dr

CHAPTER 3. THEORETICAL ANALYSIS

38

For an identical point in an arbitrary conducting body Gk :




z
dBz
2
B
=
(rA )v cos A +
(k 1)
t
dr
n
In complex form:
 



z
2
1 dBz
B
i A + 2
(k1)
i

+
(k1)
A
n
n
=
(rA )v e
+e
t
2 dr

(3.64)

This is the excitation function for the right-hand part of the governing equation (3.12).
It consists of two parts:
P art1 =

2
1 dBz
(rA )ei(A ) ei n (k1)
2 dr

(3.65)

2
1 dBz
(rA )ei(A ) ei n (k1)
(3.66)
2 dr
It is easy to see that for different ks, the components of the excitation functions
2
differ only by constant (independent of local coordinates) multipliers ei n (k1) and
2
ei n (k1) . Since we neglected the inductive properties of the conductor G ( = 0 on
the left side of (3.12)), this implies that the current densities within the conductors GK
at any instance of time consist of two components differing only by these multipliers.
After calculating the Lorenz force density and integrating over the entire volume of Gk ,
we conclude that the force acting on the conductor Gk consists of two components Fk1
and Fk2 , which are proportional to the following expressions:


i + 2
(k1)
n
(3.67)
Fk1 e

P art2 =

i + 2
(k1)
n

Fk2 e


(3.68)

Using (3.63), we find




i +2 2
(k1)

n
Fd1 e

(3.69)

Fd2 ei

(3.70)

Summing over all the conducting bodies Gk , we find that the total damping force consists of two components:

Fd1 ei

n


ei n (k1)

(3.71)

Fd2 nei

(3.72)

CHAPTER 3. THEORETICAL ANALYSIS

39

Using the equation for the sum of geometric series,


n


ei[ n (k1)] =

k=1

e4i 1
4

ei n 1
4

We limit our analysis to a particular case when n 3. In this case ei n 1 = 0, and


therefore,
n


ei n (k1) = 0

k=1

Thus, if n 3, the force exerted by the conductor G on the moving magnet is given
by
Fd = Fd2 nei

(3.73)

Combining this and the conclusion of Lemma 3, we see that the lemma is valid. In
the vector form, the force exerted by the conductor G on the moving magnet can be
presented as
F d = Cv

(3.74)

where C is the damping coefficient.

3.5

5-DOF non-contact suspension

If the magnetic field is directed axially, as it was assumed, neither of the systems described above
can produce an axially directed force. However, we know that a stable axial suspension can be
obtained easily using the interaction between permanent magnets and/or permanent magnets and
soft-magnetic components. For example, we can use the simple arrangement of permanent magnets
shown in Figure 1.1. We also know that this suspension is unstable in the radial direction, i.e.
it introduces some destabilizing negative radial stiffness, Kdes . However, utilizing the radial
suspension system described above, we can overcome this destabilizing stiffness if K cos > Kdes
and thus achieve stability in all directions.
Apparently, introducing the destabilizing stiffness Kdes in the radial plane will influence the
stability condition (3.61). However, we can easily take it into account if we redefine K and as


K

= arctan
,
(3.75)
Krad
K =


2 + K2
Krad

(3.76)

where
Krad = K cos Kdes ;

K = K sin

(3.77)

The new stability condition has exactly the same form as (3.61):
c>

sin
k
cos

(3.78)

CHAPTER 3. THEORETICAL ANALYSIS

40

Note that the final five-degree-of-freedom suspension consists of two structurally identical parts:
stationary and rotating, each including sources of magnetic fields and conductors.
Note also that in contrast to (), which varied in the range 0 < () < /2, the angle
() varies in a wider range 0 < () < . When () < /2, the net radial stiffness is
positive: Krad () = K() cos () Kdes > 0, the stabilizing radial stiffness introduced by
the first type of interaction dominates. When () > /2, the net radial stiffness is negative:
Krad () = K() cos () Kdes < 0, the destabilizing radial stiffness introduced by the second
type of interaction dominates. Stability is possible only if () < /2 (a necessary condition).
Knowing that Kdes is independent of , K() increases and () decreases with (see Lemma 3),
we conclude that () decreases monotonically with . Therefore, for a given system, saying that
the rotational speed decreases is equivalent to saying that the angle () increases and vice versa.
We will use this relationship in the following discussion when investigating the influence of the
stationary conductors inductance on the minimal stable levitation speed in this type of bearing. In
particular, we will show that the minimal stable levitation speed can be reduced through an appropriate choice of the stationary conductors inductance. Accordingly, this is equivalent to showing
that stable leviation can be achieved at higher ().
For brevity, in what follows we will not explicitly indicate the dependence of () and K()
on and simply designate them as and K unless we want to emphasize this dependence.

3.6

Effects of the stationary conductor inductance on system stability

Earlier, we showed that when the inductive properties of the stationary conductor G are neglected
and it has at least 3-rd order rotational periodicity, its interaction with a source of magnetic field
mounted on the rotor results in a purely damping force: F = Cv. In this section, we investigate
the influence of the inductive properties of the stationary conductor on system dynamics and stability. One of the interesting, practical, and important conclusions of the following analysis is that the
minimum stable levitation speed of a suspension may be reduced through an appropriate choice of
the stationary conductor inductance.
When this inductance was neglected, the necessary stability condition for the whole system in
this case was Krad > 0 ( < /2): the force acting on a rotor displaced from the central position
has to have a component pushing it back. The system was unstable if Krad 0 ( /2).
When we included the stationary coil inductances in our analysis, however, we found that the
system may be stable even if Krad 0 ( /2). The physical explanation is that, when the
stationary conductors have a non-zero inductance, they produce not only a damping force, but also
an additional positive radial stiffness.
As a simple example clarifying this point consider the system shown in Figure 3.9 consisting of three stationary coils interacting with a disk-shaped permanent magnet mounted on a rotor.
Assume that Krad 0 ( /2). When the rotor is displaced from the central position by
some distance r, a force is developed which has both radial F0r = K cos r and tangential
F0 = K sin r components. Note that when /2 < , the radial force component is
either zero or directed outwards. The tangential component causes a circular motion of the rotors
mass center around the bearing axis, which results in a centrifugal force adding to the destabilizing
effect of F0r . Thus, according to our previous analysis, the rotor center of mass would spiral away
from the bearing axis and, therefore, the system would be unstable.
The stabilizing effect of the stationary coil inductance arises because it shifts the phase of the
harmonic currents induced in the coils during the circular motion of the rotor center of mass in such

CHAPTER 3. THEORETICAL ANALYSIS

41

Figure 3.9: Explanation of the stabilizing effect due to the inductance of the stationary coils.

a manner that the force produced by these currents becomes more of a spring-type rather than a
purely damping one. In other words, an additional radial stabilizing stiffness is developed. Note
that, if the stationary coils were superconducting (zero resistance) and the currents were originally
zero when the center of the rotating magnet coincided with the bearing axis, any rotor displacement
from the central position would cause persistent currents in the coils producing a force pushing the
rotor backwards.
In general, the effect of the stationary coils cannot be reduced to pure damping whenever the
time constant of the coils (ratio of the coil inductance to its resistance) is comparable to the characteristic time of the motions of the rotor center of mass. (The coil time constant would be infinite
in the case of superconducting coils). In what follows, we do not analyze the dynamic effects of
the stationary coil inductances in detail, but instead we derive stability conditions in the presence of
this inductance using general methods of linear system theory (Routh-Hurwitz criterion).

3.6.1

Equations of motion

We assume that the stationary conductors are represented by n identical coils with inductance L and
resistance R located uniformly around the axis Z0 (Figure 3.10 shows three coils). We also assume
that the mutual inductance between the stationary coils and its effect on the system dynamics are
negligible. In practical systems employing air-coils, the mutual inductance is typically less than
the coil self inductance by at least an order of magnitude. However, we did not investigate how
sensitive to this parameter the system dynamics are. Therefore, the following analysis is based on
the assumption that the effect of the mutual inductance on the system dynamics is negligible.
The coils are exposed to a magnetic field that arises from a movable permanent magnet. The
field is circumferentially uniform about some axis Z0 . We assume that the axis Z0 is located in close
proximity to the axis Z0 . We have already established that whenever any current flows in a conductor
exposed to a circumferentially uniform magnetic field, the resulting force passes through the field
symmetry axis (Lemma 1). For simplicity, we assume that the field is directed axially. Because
of this, the Ampere forces are located in the radial plane. We introduce an in-plane Cartesian

CHAPTER 3. THEORETICAL ANALYSIS

42

Figure 3.10: Arrangement of stationary coils and a movable magnet.

coordinate system X  Y  with the origin on the Z0 axis and the X  axis directed along the line of
action of the force produced by one of the coils. Designating this coil as 1 and numbering the other
coils counterclockwise, we note that the force vector produced by the k-th coil is directed at the
angle 2(k 1)/n with respect to the X  axis. Thus the complex representation of the force vector
k = Fk ei2(k1)/n .
produced by the k-th coil is F
Assume that the magnet moves laterally with some velocity v directed at an angle with respect
to the X  axis. The complex representation of this vector is v = vei . The change of the magnetic
flux k through the k-th coil caused by the magnet motion is


2
dk
= 0 v cos
(k 1)
dt
n
We rewrite this expression as
,
,
+
2
2
2
dk
1+
= 0 vRe ei( n (k1)) = 0 v ei( n (k1)) + ei( n (k1))
dt
2

(3.79)

The current in the k-th coil is given by the following differential equation
L

dk
dIk
+ RIk =
dt
dt

(3.80)

For simplicity, we assume that the magnitude of the force produced by the k-th coil is proportional
to the current in the coil with some essentially constant proportionality coefficient I :
Fk = I Ik

(3.81)

This assumption is necessary for the system to be linear. In practice, the linearity can be achieved,
for example, if circular arc sections of the coils are exposed to magnetic fields having zero radial
gradient. Substituting (3.81) along with (3.79) into (3.80), we get

,
2
dFk
1 I + i( 2 (k1))
n
+ Fk = 0 v
e
+ ei( n (k1))
dt
2R

(3.82)

CHAPTER 3. THEORETICAL ANALYSIS

43

L
where = R
is a time constant which is a function of the coil parameters. Multiplying (3.82) by
(k1)
i2(k1)/n
k = Fk ei 2
n
e
and noting that F
we get

k
+
,
dF
(k1)
k = 0 v 1 I ei + ei ei 4
n
+F
dt
2R

(3.83)

The net force exerted by n stationary coils on the magnet is


F =

n


k
F

(3.84)

k=1

Summing equations (3.83) over all ks and recalling (3.84) produces


 4
,
1 0 I + i
dF
+ F =
v ne + ei
ei n (k1)

dt
2 R
n

(3.85)

k=1

Using the equation for the sum of geometric series,


n


ei[ n (k1)] =

k=1

e4i 1
4

ei n 1
4

We limit our analysis to a particular case when n 3. In this case ei n 1 = 0, and therefore,
n


ei n (k1) = 0

k=1

Thus, if the number of coils n 3, the force exerted by the coils on the moving magnet is given by

dF
+ F = Cvei = C
v
dt

(3.86)

where
v = vei
and
C=

n 0 I
2 R

(3.87)

If = 0, i.e.: if the inductances of the stationary coils can be neglected, then the force F is
purely dissipative and produces a damping effect: F = C
v.
Written explicitly, the force components are
dFx
dt + Fx = C x
(3.88)
dFy
dt + Fy = C y

CHAPTER 3. THEORETICAL ANALYSIS

44

We consider again the equations of motion of a portion of the suspension system including a
conductor rotating in a circumferentially uniform stationary magnetic field (3.55):

0
1
0
0
x

1
1

x 2
x2
k cos k sin k sin k cos
= [A]
; A=
.
y 1
y1
0
0
0
1

k
k

y 2
y2
cos k cos sin
k sin
Here, K and are functions of rotational speed.
If we replace the damping force in (3.55) with (3.88), we obtain the following system of equations:

x1
x1

x2
x2

d x3
x3
= [A]
;
x4
x4

dt

x
x

x6
x6

0
1
0
0
0
k cos k sin k sin k cos
1

0
0
0
1
0
.
A=

k sin k cos k cos k sin 0

0
cf
0
0
f
0
0
0
cf
0

0
0
0
1
0
f

(3.89)

where
x3 = y; x4 = y;
x5 =
x1 = x; x2 = x;

Fy K
1
C
Fx
; x6 =
; k =
; f = ; c=
m
m
m

Deriving an analytical stability condition similar to (3.61) for this system is difficult. For a
given set of parameters, the stability bounds can be found numerically for example using the simple
MathCAD script provided in the Appendix A.
Analytical stability conditions can be found however when some simplifying assumptions are
made. When analyzing (3.55), we have shown that in the presence of external damping, the speed
cross-coupling terms can be neglected and a simpler system of equations (3.60) can be analyzed
instead

0
1
0
0
x

1
1

k cos c k sin 0
x 2
x2
.
= [A]
; A=

y 1
y1
0
0
0
1

y 2
y2
0 k cos c
k sin
Similarly we may expect that the speed cross-coupling terms can be neglected when using (3.88)

CHAPTER 3. THEORETICAL ANALYSIS

45

instead of a damping force. The corresponding simplified system of equations of motion is

x1

d
x3
= [A]

dt
x4

x
5

x6

x1
x2
x3
x4
x5
x6

.
A=

0
1
0
k cos
0
0
0
k sin
0
cf
0
0

0
0
k sin
0
0
1
k cos
0
0
0
0
cf

0
1
0
0
f
0

0
0

0
.
1

0
f
(3.90)

We successfully obtained analytical stability conditions for this system. An example of a comparison of the stability bounds calculated using these conditions and exact stability bounds calculated
for the full system (3.89) can be found in the experimental section of the dissertation - the two
bounds are very close.

3.6.2

Stability conditions

The Routh-Hurwitz stability criterion requires that for the system (3.90) be stable there must be
no sign changes in the following series of coefficients calculated using the procedure described
in [110]:
a6 = 1
a5 = 2f
b5 = f 2 + cf
c5 =

2f c(f 2 + k cos )
f +c

f 2 (k cos cf )2 + f k (f + c)(f 2 cos k sin2 )


f 2 + k cos
,
+
sin2 cf 3 k k 2 f cos (c + 2f ) k (k 2 + f 4 ) + f 4 c2 cos

e5 = 2k f c
f (k cos cf )2 + k (f + c)(f 2 cos k sin2 )

d5 =

f5 = k 2 f 2
In our analysis, we assume that f > 0 ( < ) but allow cos to be negative ( > /2). Then
we have a6 > 0, a5 > 0, b5 > 0 and f5 > 0. Thus, for the system to be stable, we need c5 > 0,
d5 > 0 and e5 > 0. This leads us to three stability conditions:
f 2 + k cos > 0

f (k cos cf )2 + k (f + c)(f 2 cos k sin2 ) > 0


+
,
sin2 cf 3 k k 2 f cos (c + 2f ) k (k 2 + f 4 ) + f 4 c2 cos > 0

CHAPTER 3. THEORETICAL ANALYSIS

46

We designate
W1 = f (k cos cf )2 + k (f + c)(f 2 cos k sin2 )

(3.91)

+
,
W2 = sin2 cf 3 k k 2 f cos (c + 2f ) k (k 2 + f 4 ) + f 4 c2 cos

(3.92)

and

Then the stability conditions become

3.6.3

f 2 + k cos > 0

(3.93)

W1 > 0

(3.94)

W2 > 0

(3.95)

Analysis of the stability conditions

We analyze the stability conditions (3.93), (3.94) and (3.95) for three separate cases: cos > 0,
cos = 0 and cos < 0.

Case 1: cos > 0

( < /2) .

If the inductances of the stationary coils were neglected ( 0 or f ), then in this case the
stability conditions would reduce to the single condition (3.61). To show this, we first note that the
condition (3.93) is clearly satisfied in this case. Next, we expand W1 and W2 in powers of f :
W1 = (c2 + k cos )f 3 k cos cf 2 + (k 2 cos2 k 2 sin2 )f ck 2 sin2 (3.96)
W2 =(k sin2 + c2 cos )f 4 + (ck sin2 )f 3 (2k 2 sin2 cos )f 2
(ck 2 sin2 cos )f k 3 sin2

(3.97)

Looking at the signs of the higher order terms, we conclude that when f is large ( 0) , the
first condition W1 > 0 is always satisfied, while the second condition is satisfied only if
k sin2 + c2 cos > 0
We rewrite this condition as
c>

sin
k
cos

this is the condition (3.78).


Further, we will prove the more general fact that (3.94) is always satisfied if (3.95) is satisfied,
and therefore only (3.95) must be considered. For this purpose, we expand W1 and W2 in powers
of c:
W1 = f 3 c2 k [f 2 cos + k sin2 ]c + [k 2 f cos2 + k f (f 2 cos k sin2 )] (3.98)

CHAPTER 3. THEORETICAL ANALYSIS

47

W2 = [f 4 cos ]c2 + [f k sin2 (f 2 k cos )]c k sin2 (k 2 + 2k f 2 cos + f 4 ) (3.99)


We form a new function W1 by multiplying W1 by a positive number f cos : W1 = W1 f cos .
Clearly, the condition that W1 > 0 is equivalent to W1 > 0. Note that the difference W = W2 W1
is a monotonically increasing function of c:
+
,
(3.100)
W = k f 3 c (k 2 sin2 + f 2 (f 2 + k cos ))
The functions W2 and W1 become equal at some positive value c0 of c:
c0 =

k 2 sin2 + f 2 (f 2 + k cos )
f3

(3.101)

For all c > c0 , W > 0, i.e.: W2 > W1 . For all c < c0 , W < 0, i.e.: W2 < W1 .
At the point c = c0 where W1 = W2 , we define the positive value of W1 and W2 :
.
W0 = W1 (c0 ) = W2 (c0 ) = f 2 cos (f 2 + k cos )2

(3.102)

The derivative of W1 with respect to c is a monotonically increasing function of c:


dW1
= 2f 4 cos c f k [cos2 (f 2 k cos ) + k cos ]
dc

(3.103)

At c = c0 this derivative has a positive value:


dW1 -= f k 2 cos sin2 + f 3 cos (k cos + 2f 2 )
dc c=c0

(3.104)

Therefore, function W1 is positive at c = c0 and at all c > c0 . Since we also found that for
c > c0 , we have W2 > W1 , both stability conditions (3.94) and (3.95) are satisfied when c c0 .
For c < c0 , we have W2 < W1 . If we require W2 > 0, then W1 > W2 > 0 for c < c0 and stability
conditions (3.94) and (3.95) are satisfied in this range of c as well.
Thus we conclude that condition (3.95) is necessary and sufficient for the systems stability.
Defining = 1/f produces a more convenient form of (3.95):
cos c2 + k sin2 (1 k 2 cos )c k sin2 (1 + 2k 2 cos + k 2 4 ) > 0 (3.105)
Solving this inequality for c, we find the following lower bound on the damping coefficient
needed for system stability:
k sin2 (k 2 cos 1) +
c>
2 cos

(3.106)

where D is the determinant of the left-hand part of the equation (3.105), which can be written as

D = k cos sin k 2 +

1
cos

2 

4 cos
k +
sin2
2


(3.107)

CHAPTER 3. THEORETICAL ANALYSIS

48

We introduce new dimensionless variables


u = k 2 ;

u1 =

4 cos
;
sin2

u2 =

1
cos

(3.108)

The meaning of these designations will become more transparent when we analyze the case >
/2.
In terms of the new variables, the stability condition becomes

u2 u
u2 + u
c > c1 = 2
k u+2
k (u + u1 )
(3.109)
u2 u1
u2 u1


1It is easy to see that when u = 0, c1 -u=0 = 2 uk1 > 0; dc


du u=0 = and c1 u +.
Therefore, there is an optimal value of u resulting in minimal c1 .

Case 2: cos = 0 ( = /2) .


The system was clearly unstable in this case when we neglected the inductive properties of
the stationary coils since the net radial suspension stiffness was zero and the centrifugal force was
pushing the rotor away from equilibrium. Equation (3.78) requires that infinite damping is needed
for stability when cos +0 (this equation is not applicable for cos < 0). However, reviewing
(3.95) and (3.92), it is clear that this stability condition can be satisfied with = /2 if cf 3 >
(k 2 + f 4 ). This obviously requires that the coils time constant, = 1/f , be a non-zero . Further,
we will show that as in the first case, (3.94) is always satisfied when (3.95) is satisfied. The functions
W1 (3.98) and W2 (3.99) in this case simplify to
W1 = f 3 c2 k 2 c k 2 f

(3.110)

W2 = f 3 k c k (k 2 + f 4 )

(3.111)

The first function is positive if



k 2 + k k 2 + 4f 4
c > c01 =
2f 3

(3.112)

while the second function is positive if


c > c02 =

k 2 + f 4
f3

(3.113)

We will show that c02 > c01 and, therefore, if W2 > 0, then W1 > 0. We form a difference:
= 2f 3 (c02 c01 )
The original statement that c02 > c01 is equivalent to requiring > 0. Writing explicitly:

= 2(k 2 + f 4 ) k 2 k k 2 + 4f 4
We rewrite this expression as
=ab

.
a = k 2 + 2f 4 > 0 and

. 
b = k k 2 + 4f 4 > 0

CHAPTER 3. THEORETICAL ANALYSIS

49

Saying that c02 > c01 is equivalent to saying that a > b. Squaring both a and b, we get
a2 = k 4 + 8k 2 f 4 + 4f 8

(3.114)

b2 = k 4 + 4k 2 f 4

(3.115)

Comparing (3.114) and (3.115), one can see that, indeed, a > b. Thus, in this case we again need
only one condition to be satisfied for the system to be stable:
c>

(k 2 + f 4 )
f3

This condition is also more convenient to use if expressed in terms of :


c>

k 2 4 + 1

or in terms of u:
.
c > c1 =

(3.116)

k 2
(u + 1)
u

Case 3: cos < 0

(3.117)

( > /2) .

Here, we must consider the condition (3.93) along with (3.94) and (3.95), since when cos < 0
this condition can be violated. Again, we will show that the condition (3.95) is necessary and
sufficient for system stability. We will also derive a useful necessary stability condition that gives
us a more convenient way of finding a range of where stability can be achieved, rather than by
analysis of (3.95). For each from this range, there exists a certain range of c, where the system is
stable. This range of c then can be found from (3.95).
As a first step, we rewrite (3.93) for this case as
f 2 > k | cos |

(3.118)

Further, as in case 1, we consider a function W1 defined as W1 = (f cos )W1 . This time, however,
cos < 0, and, therefore, in terms of W1 , (3.94) becomes:
W1 < 0

(3.119)

Note that both equations W1 (c) = 0 and W2 (c) = 0 describe parabolas with branches extending
downwards and the extrema being the function maxima in contrast to case 1, since the coefficients
of the higher order term are now negative. Thus, the condition W2 > 0 can be satisfied in the area
between two values of damping c1 and c2 which are solutions of the equation W2 = 0. The system
becomes unstable if there is either not enough damping (c < c1 ) or if there is excessive damping
(c > c2 ).
As in case 1, the functions W2 and W1 become equal at some value c0 of c (see equation (3.101)).
.
If the condition (3.118) is satisfied, the value c0 is clearly positive. The value W0 = W1 (c0 ) =
W2 (c0 ), is a negative number (see equation (3.102)). The derivative of W1 with respect to c is a
monotonically decreasing function of c (see equation 3.103). At c = c0 this derivative is negative

CHAPTER 3. THEORETICAL ANALYSIS

50

(equation 3.104). Therefore, it is negative for all c > c0 and W1 is also negative in this range. Thus,
the condition (3.119) is satisfied for all c > c0 .
The functions W2 (c) and W1 (c) can cross each other on either side of the maximum of W2 (c).
More specifically, since we established that the function value at the crossing point is negative
(W0 < 0), the crossing point is either less than c1 (c0 < c1 ) or greater than c2 (c0 > c2 ). The
first case is represented by the curves W2 and A on Figure 3.11. The second case is represented by
the curves W2 and B. It is easy to see that in the first case, both conditions W2 > 0 and W1 < 0
B

W2
0
C2

C1

Figure 3.11: Possible mutual orientations of the graphs W2 (x) and W1 (x).
are satisfied in the damping range c1 < c < c2 . In the second case, these conditions cannot be
satisfied simultaneously. Indeed, equation (3.100) shows that the difference W = W2 W1 is a
monotonically increasing function of c. For c < c0 , we have W < 0, i.e. W2 < W1 . Therefore,
if c0 > c2 , then in the interval c1 < c < c2 , we have 0 < W2 < W1 and only one of the stability
conditions is satisfied (we need W1 < 0, see (3.119)).
Thus, instead of (3.119), we can use the following: the system is stable if the curves W2 (c) and
W1 (c) cross each other on the left from the maximum of W2 (c) and is unstable otherwise. This
2
condition is equivalent to requiring that the derivative dW
dc is positive at c = c0 . The latter is given
by the following expression:


dW2 -(3.120)
= f 2f 4 cos + f 2 k (1 + cos2 ) + k 2 cos sin2
dc c=c0
Therefore, the condition (3.119) is equivalent to the following:
2f 4 cos + f 2 k (1 + cos2 ) + k 2 cos sin2 > 0

(3.121)

Solving for the roots of the left side of (3.121), we find two solutions:
f12 = k cos = k | cos |
and
f22 = k

2
sin2
sin
=
k
2 cos
2| cos |

Considering that the coefficient in front of the highest power of f in condition (3.121) is negative,
it is satisfied if
k | cos | < f 2 < k

sin2
2| cos |

(3.122)

CHAPTER 3. THEORETICAL ANALYSIS

51

Note that the left inequality simply repeats the condition (3.118). Therefore, if (3.122) is satisfied,
then both (3.118) and (3.94) are satisfied. Thus, at this point we conclude that two conditions (3.122)
and (3.95) define the systems stability.
The condition (3.122) gives us some range of where stability can be achieved:
1
2| cos |
2
2 < < k | cos |

k sin

(3.123)

The condition (3.95), when expressed in terms of , has the same form as (3.105) in case 1.
Similar to cases 1 and 2, this condition imposes bounds on the amount of damping, given a certain
value of . However, in contrast to those cases, now there are both lower and the upper bounds
(because the coefficient in front of c2 is negative):
k sin2 (k 2 | cos | + 1)
2| cos |

k sin2 (k 2 | cos | + 1) +
<c<
2| cos |

(3.124)

where

D = k cos sin k 2

1
| cos |

2 

4| cos |
k
sin2
2


(3.125)

(This is Eq. (3.107), adapted for the case > /2 by replacing cos with | cos |).
As before, we rewrite this conditions in terms of dimensionless variables u, u1 and u2 . This
time, however, considering that cos is negative we redefine u1 and u2 as
u1 =

4| cos |
;
sin2

u2 =

1
;
| cos |

(3.126)

It is easy to see that the condition (3.123) in terms of the u variables becomes
u1
< u < u2
2

(3.127)

Thus, u1 and u2 relate to upper and lower bounds on u = k 2 .


Using u variables, we rewrite (3.125) as
D = k cos2 sin4 (u u2 )2 (u u1 )

(3.128)

Note that the existence of a solution requires D 0, which is possible if and only if u u1 . This is
a stricter lower bound on u than given by (3.127). Ignoring for practical reasons the boundary value
u = u1 , we state that the system can be stable only if (necessary condition)
u1 < u < u2

(3.129)

or in terms of :
1
4| cos |
2
2 < < k | cos |

k sin

(3.130)

CHAPTER 3. THEORETICAL ANALYSIS

52

For each , from the range (3.130) (or, identically, for each u = k 2 from the range (3.129)), there
exists a range of c given by (3.124), where the system is stable. In terms of u, (3.124) becomes:

(
u2 + u
u2 u
k u2
k (u u1 ) < c ;
c1 = 2
u2 u1
u2 u1

(
'
u2 + u
u2 u
k u+2
k (u u1 )
c < c2 = 2
u2 u1
u2 u1

'

(3.131)

It is easy to see that satisfaction of (3.131) in fact implies that (3.129) is satisfied. Indeed, the
roots c1 and c2 in (3.131) exist if and only if u > u1 . Further, if the roots exist, then c2 > c1 if and
only if u2 > u. Thus satisfaction of (3.131) implies that (3.129)) is satisfied, and, therefore, (3.131)
is proven to be necessary and sufficient for the system stability.
The necessary stability condition (3.129) is very useful in practice because it directly gives us a
range of where stability could be achieved through appropriate choice of c. The range of c for a
given can then be found using (3.131).
Note that values of , which make the system stability possible, exist if and only if
1
4| cos |
2 < k | cos |

k sin
or
tan > 2

(3.132)

(or, equivalently, the lower bound on the stable levitation


This gives us the upper bound on , max

speed min : max = (min )):

= arctan(2) = 0.648 = 2.04 rad


max

(3.133)

This value of min is lower than the one which could be obtained with = 0 (the latter corresponds
to = /2). Note, that the lower bound on the minimal stable levitation speed does not depend on

we must choose
the in-plane stiffness k . To approach = max
u = uopt =

4| cos max
|

sin2 max

| cos max
|

2.236

(3.134)

The results of the stability analysis are summarized in the Table 3.1.
To illustrate the above conclusions, we have plotted the bounds
on the amount of damping
needed for system stability (represented by dimensionless s = c/ k , see Eq. (3.109), Eq. (3.117)
and Eq. (3.131)) as a function of the angle . Fig. 3.12 shows several plots obtained for different
time constants of the stationary coils, represented by dimensionless u = k 2 .
One can see that if u = 0, the amount of damping s goes to infinity when approaches /2.
Stability is not possible beyond the line = /2. With u = 0.5, however, the boundary of the
stability region penetrates into the earlier prohibited area > /2. Also note that with u = 0.5
less damping is needed for stability at < /2. With a further increase of u up to 1 and 2.236,
the stability region penetrates deeper on the right of the line = /2, however more damping is
needed for the system to be stable at < /2. Note that the boundaries of the stability regions
in the area > /2 are nose-shaped: there are both upper and lower bounds on the amount of

CHAPTER 3. THEORETICAL ANALYSIS

u=0
( = 0)
u>0
( > 0)

< /2
c > c1 |u=0 (a)
(Eq. (3.109))
or Eq. (3.78)
c > c1 (u)(a)
(Eq. (3.109))

53
= /2

> /2

Unstable

Unstable

c > c1 (u)(a)
(Eq. (3.117))

u1 < u < u2 (b)


c1 (u) < c < c2 (u)(a)
(Eq. (3.131))

Table 3.1: Summary of the stability conditions.


Designations: u = k 2 ; u1 = 4| cos | csc2 ; u2 = | sec |
(a)The condition is necessary and sufficient
(b)The condition is necessary but not sufficient

damping. In the area < /2, there is only a lower bound. The value of u = 2.236 corresponds
to the upper bound on (see Eq. (3.133) and Eq. (3.134)).
If we continue increasing u, the shapes of the curves change as can be observed for u = 3: the
curves c2 ( ) and c1 ( ) cross each other before reaching the point when c2 and c1 do not exist
(D < 0 in Eq. (3.128)). Let us analyze this feature in more detail. Note that at any crossing point
where only one root exists, D = 0. This is possible if and only if u is equal to either of its extreme
values: u1 or u2 (see Eq. (3.128)). Also note, that for > /2, u1 increases monotonically and u2
decreases monotonically with (see Eq. (3.126)), as indicated in Figure 3.13.
If u < uopt , with increasing first we get u = u1 and only then u = u2 . However, after the
point u = u1 we have u < u1 and, consequently, D < 0 (see Eq. (3.128)). This implies that the
roots c1 and c2 in Eq. (3.131) do not exist. The point when u = u1 is the only point when we have
c 2 = c1 .
If u > uopt , with increasing first we get u = u2 and then u = u1 . In this case after the point
u = u2 (where c2 = c1 ) we still have D > 0 (Eq. (3.128)) and the roots c1 and c2 in Eq. (3.131)
exist until the point u = u1 . This gives us the part of the curve plotted for u = 3 on the right of
the first crossing point on Fig. 3.12. However, after the point u = u2 the condition Eq. (3.129) is
violated since u > u2 . Moreover, the condition Eq. (3.131) is violated as well because u > u2
results in c1 > c2 .

3.7

Rotational loss due to the rotor unbalance

So far we considered an ideal situation assuming that certain axes of the system coincide. In the
Table 3.2 we list the important axes, which define the system behavior.
All the axes are assumed to be collinear.
Note, that we introduce three new axes: Zar , Zas and Zm . Axes Zar and Zas define equilibrium
of the axial suspension in the radial plane: when these axes coincide, the axial suspension produces
no force in the radial plane. The axis Zm is the axis on the rotor passing through its center of mass.
Earlier we assumed that the axes Zar and Zm coincided with the axis Z0 and the axis Zas coincided
with Z0 . In principle, we could also introduce a designation for the symmetry axis of the rotating
conductor G, however, in what follows we assume that there is no radial loading, in which case the
shape of the conductor G, including location of its symmetry axis, appears not to be important (see

CHAPTER 3. THEORETICAL ANALYSIS

10

54

u=0

8
6

u=3

4
2

u=2.236
u=1

u=0.5
0
0.4

0.5

0.6

0.7

Figure 3.12: Stability regions calculated for different u values. s = c/ k .

Figure 3.13: Analysis of the stability region plotted for u = 3 (see Figure 3.12).

for example the discussion in Section 2.7 and in the following section).
An analysis of the system when all the axes are different is going to be complicated. Here we
consider a simpler case: we consider rotational losses in the radial part of the bearing caused by a
misalignment of the axes Zm (rotor center of mass) and Z0 (rotor magnetic field symmetry axis).
The analysis illustrates how different parameters of the radial suspension affect the rotational loss in
this case. For simplicity, we also assume that the stationary conductor symmetry axis Z  coincides
with the stationary field symmetry axis Z0 . Since we do not include the axial suspension, we do not
need axes Zar and Zas for this analysis.

3.7.1

Estimation of the rotational loss in the radial suspension neglecting the


stationary conductor inductances

Earlier we found that the electromagnetic force acting on a conductor rotating in a circumferentially
uniform magnetic field is zero whenever its rotation axis Z coincides with the field symmetry axis
Z0 . Note that the rotation axis Z can be any axis of the rotor parallel to Z0 . This situation is
very different from that for Active Magnetic Bearings in their basic form, conventional Null-Flux

CHAPTER 3. THEORETICAL ANALYSIS


Z
Z
Z0
Z0
Zar
Zas
Zm

55

Rotor rotation axis;


Stationary conductor symmetry axis;
Stator magnetic field symmetry axis;
Rotor magnetic field symmetry axis;
Axial suspension axis on the rotor;
Axial suspension axis on the stator;
Axis on the rotor passing through its center of mass.
Table 3.2: Important axes of the system.

Bearings and most other magnetic bearings where the rotor geometry dictates about which rotor
axis the rotation should take place in order for the net electromagnetic force to be zero. Dynamic
equilibrium of the rotor, however, is possible for the system presented here only if the rotation axis
Z passes through the mass center: otherwise the rotor will be influenced by the centrifugal force
which is not counteracted by the electromagnetic force.
We also found that the inverse system including a source of magnetic field mounted on the
rotor and circumferentially uniform about some axis Z0 (Q ), and a stationary conductor G can
be used to provide damping needed for system stability. (Strictly speaking the inverse system
acts as a damper only if the inductive properties of G are neglected as discussed in the previous
section). An interesting feature of this damper is that it does not affect the rotation about axis Z0 ,
but effectively damps out the lateral motions of this axis. As a result, the rotation always takes
place around the mass center provided that the axis Z coincides with the axis Z0 and they both
pass through the rotor mass center. As discussed briefly in Section 2.7, this is necessary to uniquely
define equilibrium in this system.
What will happen, however, if the axes Z and Z0 do not pass through the mass center? It is
easy to see that in fact we need to worry only about the Z0 axis, since the location of the Z axis is
entirely determined by the system dynamics: if the axis Z0 passes through the mass center, then the
axis Z will coincide with Z0 . The geometry of the conductor G is of no importance in this case, the
only important fact is if the axis Z0 passes through the mass center or not. In the case where the axis
Z0 does not pass through the mass center, we can no longer expect that either the mass center lies
on the axis Z0 or the axes Z and Z0 coincide with Z0 . However, because of the system symmetry,
in the state of dynamic equilibrium, the axes Z, Z0 and the rotor mass center must move around the
axis Z0 in circles of constant radii.
Here we make one more interesting observation about the rotation of the conductor G in a
circumferentially uniform magnetic field. Consider a particular case where the axis Z moves around
the axis Z0 in a circle of constant radius r0 with rotational speed equal to the rotational speed
around the axis Z, . (Note that this case does not qualify as slow motion, see conditions of slow
z
B
motion (3.45).) Substituting x0 = r0 cos t and y0 = r0 sin t in (3.40), we see that t
0.
Therefore, motion of the axis Z around the axis Z0 in a circle of a constant radius with rotational
speed does not cause any electromagnetic force. This is not surprising, because the composite
rotation consisting of the rotation of the body G about the axis Z with angular speed and the
rotation of the axis Z about the axis Z0 with the same speed is simply equivalent to the rotation of
the body G about the axis Z0 . Therefore, without loss of generality we can say that the rotation of
the body G still takes place around the axis Z0 . As we already know, this rotation does not produce
any electromagnetic force. The only non-zero forces which we have left are the centrifugal force Fc

CHAPTER 3. THEORETICAL ANALYSIS

56

and the damping force Fd . The dynamic geometrical sum of these forces must be zero in order to
avoid lateral motions of the rotor (the sum of the moments, however, does not have to be zero: it is
equal to the drag torque).
Let the distance from the axis Z0 to the axis Z0 be rd and the distance from the axis Z0 to the
rotor mass center be rm . These distances are yet to be determined. The force diagram is shown in
Figure 3.14. The origin of the auxiliary coordinate system Xm Ym is on the axis Z0 and the axis

Figure 3.14: Force diagram with rotor unbalance. Stationary coil inductances are neglected.
Xm passes through the mass center of the rotor. This coordinate frame rotates with angular speed
around the axis Z0 . The centrifugal force Fc is clearly directed along the Xm axis. Using Lemma
4, we note that the direction of the damping force Fd is perpendicular to the vector rd drawn from
the axis Z0 to the application point of this force. Clearly, the forces can balance each other in the
direction of the Ym axis if and only if the vector rd is perpendicular to the vector rm drawn from
the axis Z0 to the mass center (Xm axis). (In the diagram shown in Figure 3.14 we also used the
fact that the application point of the damping force is always on the Z0 axis (Lemma 1).)
From the force balance in the direction of the axis Xm , we obtain
m 2 rm = Crd

(3.135)

Using orthogonality of the vectors rm and rd, we get


2
+ rd2
2u = rm

(3.136)

Note that since u is the parameter which characterizes the unbalance, it is equal to the distance
from the rotor mass center to the field symmetry axis Z0 . Solving the equations (3.135) and (3.136)
together, we find
r d = u

m
+ C2

(3.137)

m2 2

Then the drag torque can be written as


Tdrag = Crd2 = C2u

m2 3
m2 2 + C 2

(3.138)

and the rotational loss becomes


Ploss = Tdrag = C2u

m2 4
m2 2 + C 2

(3.139)

CHAPTER 3. THEORETICAL ANALYSIS

57

In practice, usually C 2 << m2 2 , in which case the rotational loss can be estimated as
Ploss = C2u 2

(3.140)

Note that this loss is essentially independent of the rotor mass. From the equation (3.135) we see
that if C << m2 then rm << rd u , i.e. the rotation takes place about the mass center.

3.7.2

Estimation of the rotational loss in the radial suspension considering the


stationary conductor inductances

It can be easily shown that when the stationary conductors have inductive properties, the vector of
the force Fd acting on the field source Q during a circular motion of its axis Z0 around the axis
Z0 will lag the vector opposing the velocity of Z0 by some angle 0 < < /2 (we use the same
symbol Fd as before, even though this force is not a damping force anymore). If the stationary
conductor is represented by a set of separate coils with inductance L and resistance R, then
tan =

L
.
R

(3.141)

Also, because of the overall increase of the stationary conductor impedance, the magnitude of Fd
+
,0.5
will be lower than when the inductances were neglected by the factor 1 + tan2
, i.e.
C
Fd = 
rd
1 + tan2

(3.142)

The force diagram corresponding to the dynamic equilibrium in case when the stationary conductor inductances are taken into account is shown in Figure 3.15. In this case instead of the equa-

Figure 3.15: Force diagram with rotor unbalance. Stationary coil inductances included.

tion (3.135) we have


C
rd
m 2 rm = 
1 + tan2

(3.143)

and instead of the equation (3.136) we have


2
2
+ rd2 2rm rd cos(/2 ) = rm
+ rd2 2rm rd sin
2u = rm

(3.144)

CHAPTER 3. THEORETICAL ANALYSIS

58

From (3.143) and (3.144) we find


rd = u ,

(3.145)

where

= 1 +

1
C

m 1 + tan2

0.5

2
2

sin
C


m 1 + tan2

(3.146)

C
C
rd2 cos = 
2 2u cos .
Tdrag = 
2
1 + tan
1 + tan2

(3.147)

The drag torque is

,0.5
+
, we rewrite (3.147) as
Using the trigonometric identity cos = 1 + tan2
Tdrag =

C
2 2u .
1 + tan2

(3.148)

Note, that the inclusion of the stationary conductor inductance reduces the drag torque not only
because the magnitude of Fd is reduced but also because the lever arm of Fd becomes smaller and
approaches zero when /2.
The rotational loss can be found as
Ploss = Tdrag =

C
2 2 2u .
1 + tan2

(3.149)

If the assumption C << m is justified, then of course we will also have C << m
In this case 1 and rd u . Then the equation for the rotational loss becomes
Ploss

1 + tan2 .

C
2 2u .
1 + tan2

Note, that as before, Ploss is independent of the rotor mass. Assuming (tan )2 =
we rewrite (3.150) as
Ploss

CR2 2

L2 u

(3.150)
 L 2
R

>> 1,

(3.151)

when the inductive component of the stationary coil impedance dominates, the rotational loss due
to the unbalance is independent of the rotational speed .
Note, that the stability conditions derived earlier in this dissertation (Eq. (3.109), Eq. (3.117)
and Eq. (3.131)) are based on the assumption that the rotor mass center coincides with the rotor field
symmetry axis. If this is not the case, the stability conditions may be different. This aspect is not
investigated in the present dissertation and may be suggested as a subject for future research.

Chapter 4

Technical and Practical Considerations


4.1

Angular stability of the suspension

In the analysis presented in the previous chapter, we considered only the systems stability with
respect to lateral motions: axial and radial. We did not consider the stability with respect to the
angular deflections about the horizontal axes. In this section, we show that this kind of stability can
be achieved rather easily through appropriate design. (Note that the Earnshaws theorem considers
only axial motion: the equation (1.3) does not involve the angular coordinates.)
Consider a simple example clarifying this point. Assume that we want to compensate for the
weight of a bar using forces of a magnetic nature (Figure 4.1). First, we attempt to do so by locating a magnet in the middle of the bar, which is repelled by another stationary magnet located
below. This pair of magnets produces a lifting force F1 (Figure 4.1). If the magnets are not in
horizontal equilibrium, this force has both vertical and horizontal components: F1z and F1y . Earnshaws theorem states that this system if stable in the vertical direction, is unstable horizontally.
1z
In mathematical terms, if the vertical stiffness is positive: K1z = F
z > 0, then the horizonF
tal stiffness is negative: K1y = y1y < 0. In our bearing, we have a rotor instead of the bar,
and the negative horizontal stiffness is compensated using interaction of the conductors mounted
on the rotor with stationary magnetic field. Clearly, the system shown in (Figure 4.1) does not
produce much of the angular stiffness about the X axis. Assume that we added two other magnet
couples at the ends of the bar, which produce positive vertical stiffness K2z > 0. Also assume that
K2z << K1z and |K2y | << |K1y |, i.e. these magnets do not affect much the overall stiffness of the
suspension. If the bar is deflected by some angle about the X axis, this will produce a restoring
torque Tx = 4K2z l2 . Thus, the angular stiffness due to two magnets on the edges of the bar is

Figure 4.1: An explanation of static angular stabilization

59

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

60

Figure 4.2: An explanation of stabilizing mechanism due to the gyroscopic effect


Kang = 4K2z l2 . By choosing large l we can obtain large angular stiffness with small K2z , and ,
therefore, ensure angular stability without affecting lateral stability.
In the case of a rotational bearing with a disk-shaped rotor, the problem of angular stabilization
is further facilitated by gyroscopic effects. Consider a disk shaped rotor (Figure 4.2) with polar
moment of inertia JP and transverse moments of inertia JT , rotating around its polar axis with
spin speed . If we apply a disturbance torque T around one of the transverse axes, it causes a
precession around the perpendicular transverse axis with precession speed
T
H

p =

(4.1)

where H = Jp and it is assumed that JT << H. The deflection of the rotor axis due to the
disturbance torque can be found by integration of (4.1):

1
p dt =
H

T dt

The higher H, the bigger the disturbance


exceeds a limiting value.

4.2

(4.2)
T
0

T dt the system can tolerate before the rotor deflection

Designing the rotating conductors and stationary magnetic field

While Figure 1.2 indicates the major components of the suspension, its performance can be improved dramatically through careful design of the conductors and the fields. The major goals encountered at this stage are minimizing the angle (which causes reduction of the minimal levitation
speed), maximizing the suspension stiffness, minimizing rotational loss and maximizing the rotor
strength. The measures leading to these goals often contradict each other and, therefore, the design
is a tradeoff between the different criteria.
One part of this design is choosing the shapes of the rotating conductors and the stationary
magnetic field. Here we will use the ideas developed earlier for Null-Flux Bearings [74, 75, 85]
and magnetic bearings using Hard Type II superconductors [55, 56]. From studies of these bearings
we see that a particular topology when a conductor is shaped as a planar loop with an opening in

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

61

Figure 4.3: A motivation for the choice of the conducting loop and magnetic field shapes.

Figure 4.4: A particular case of the general structure shown in Figure 3.1.
the center (as in contrast to a continuous slab) and exposed to a specifically shaped magnetic field
normal to the loop plane is beneficial for maximizing the load capacity and stiffness, as well as
minimizing energy dissipation when the system is out of equilibrium.
To understand this, we consider a planar conducting loop of rectangular shape, two opposing
sides of which (aligned along the Y axis) are exposed to two strongly different magnetic fields B1
and B2 normal to the loops plane (Figure 4.3). If we require that the fields B1 and B2 remain
constant for the whole range of the loop displacements in the X direction, then the field variation
during a displacement will take place only on the sides parallel to the displacement. In other words,
by removing the central part of the loop we reduce the portion of the conductor volume experiencing
field variations.
In the case of hard superconductors, this results in a significant reduction of the force-displacement
hysteresis directly relating to the remagnetization hysteresis of a hard superconductor (the volume
of the superconductor subjected to the remagnetization is reduced). In case of Null-Flux Bearings
this reduces the energy losses due to eddy currents induced in a moving conductor.
To make use of this topology in our bearing, we shape the conductor G as several closed loops
exposed to a magnetic field as shown in Figure 4.4. According to the result above, we need at least
three loops (the number of the loops represents the rotational periodicity of the conductor G) to
have a time-invariant force response to a fixed rotor displacement. This is a particular example of
the system shown in Figure 3.1. By reducing the conductor volume exposed to varying field we

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

62

Figure 4.5: Cross-section of the loop and magnetic field as in Figure 4.3 in the XZ plane.

Figure 4.6: A system XZ-plane cross-section with a conducting slab instead of the conducting loop
shown in Figure 4.5
apparently reduce unwanted eddy currents induced during the rotor rotation when it is displaced
from the equilibrium position and, correspondingly, the rotational drag.
Another advantage of using conducting loops with the central parts removed instead of solid
conductors is increase of the suspension stiffness and load capacity. To clarify this point, assume
for simplicity that Bz2 = Bz1 = B0 and the profile of the magnetic field B(x) can be reasonably
approximated as

B0 if x < b
B0 xb if b x b .
B=
(4.3)

B0
if x > b
This profile and the location of the conducting loop with respect to the profile are indicated in
Figure 4.5. Now, assume that instead of the conducting loop as in Figure 4.5 we use a solid slab
without the central opening (Figure 4.6).
Note that in both cases when either the loop or the slab moves to the right, the magnetic field
changes only in the area [b; b]. Moving the conductor to the right in the stationary magnetic field
causes the same field variation within the conductor as shifting the field to the left with the conductor

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

63

being stationary. This is schematically illustrated in Figure 4.6. For small displacements X, the
profile of the field variation can be described by a rectangle with a width of 2b and a height of
B = B0 X
b :

B if |x| b
B=
.
(4.4)
0
if |x| > b
For simplicity, in both cases we assume that the resistance of the conductor is zero or the inductive component of the impedance dominates over the resistive. Then, when the loop or the slab is
displaced in the X direction, currents develop in the conductor so as to compensate for changes of
magnetic fluxes through any closed loops located fully in the conducting volume. This can be better
understood if we look at the field diffusion equation (3.11).
Assuming that is infinite leads us to

Hj

= B
z (r, r0 , t)ez ,
t
t

or
z (r, r0 , t)ez = constant.
Hj + B

(4.5)

Similarly, if Bz (t) is a harmonic function with frequency , then Hj (t) would be a harmonic
function as well, and Bz /t as well as Hj /t . When is high, the latter term will
dominate over the spatial derivative on the left side of the field diffusion equation (3.11) which,
again, degenerates to (4.5).
Note that in the case of an infinitely long conducting slab, cylinder or, more generally, any object
with ellipsoidal cross-section which experiences a variation of the external field applied parallel to
the objects long side, (4.5) is satisfied if the current flows in an infinitely thin layer adjacent the
object surface. This result can be also obtained if we recall the familiar equation for the skin depth:
.
2
.
=

If either or go to infinity, the skin depth goes to zero.


In contrast, the conducting slab shown in Figure 4.6 experiences a field variation which is perpendicular to the slab. It is easy to see that since the slab is of finite length in the field (Z) direction,
(4.5) cannot be satisfied if the currents flow only on the edges of the field variation zone. To reproduce the square shape of the external field variation, they must flow in the area [b; b] . The skin
effect will manifest itself in pushing currents on the slab edges in the Z direction. If we assume,
however, that the conductivity is infinite, then the current value does not depend on how thin the
layer where the current flows is: the path resistance is zero anyway. Without this assumption the
averaged over the slab thickness value of the current at any frequency would be lower. Thus, by
assuming = we give a slab conductor an advantage in our comparison with the conducting
loop.
As we mentioned, in the case of a conducting slab shown in Figure 4.6, the compensation of the
external field variation would be in large degree achieved by the currents flowing in the area of the
non-uniform field: [b; b], where the absolute value of the field is lower than its maximum (B0 ).
In contrast, if the central part of the conducting slab is removed (conducting coil case), the currents
are forced to flow in the areas where the magnitude of the field is maximal (B0 ). This results in

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

64

a higher stiffness and load capacity. Furthermore, to reduce and essentially to eliminate the skin
effect, the conducting loop can be wound of several turns of insulated wire (the wire ends will need
to be connected to form a closed loop). It will be shown later in this chapter that making the loop
out of wire does not significantly effect the loading characteristics of this system provided that the
copper factor is close to unity. On the other side, using wire results in uniform distribution of the
current in the loop cross-section which is guaranteed by the current continuity along the wire. The
skin effect in this case is essentially eliminated if the wire radius is less than the skin depth.
We can make a rough estimate of how much bigger would be the stiffness in case of the conducting loop as compared to the zero-resistivity slab. First consider the case of the conducting slab
(Figure 4.6). To calculate the force acting on the slab, we need to solve an inverse problem and find
the current distribution which results in the magnetic field given by (4.4). We assume that the slab
is somewhat wider than 2b and has a width of 2b where > 1. We also assume that the slab is
thin, i.e. it has the thickness t << 2b. Considering the problem symmetry we look for the current
per unit length of the slab in the X direction, J, in form of an expansion over x limiting it to the first
two odd-order terms:
J(x) = 1 x + 2 x3 .

(4.6)

In case of a thin slab, the magnetic field in the middle of the slab can be found as
0
B(x) =
2

b
b

J(x0 )
dx0 .
x0 x

(4.7)

Instead of the coefficients 1 and 2 in (4.6) we introduce additional coefficients and so that
1 =

;
b

2 =

(b)3

(4.8)

The coefficient allows us to scale current and field distributions ((4.6) and (4.7)) without changing
their shapes, while the coefficient affects the shapes. It is convenient to describe the field and
current distributions in terms of the dimensionless variable x
= x/b. Then, the current per unit
length is given by

 
x

J(
x) =
+
.
(4.9)

and the equation (4.7) becomes


B(x) =

0
2

J(
x0 )
d
x0 .
x
0 x

(4.10)

Evaluating (4.10), we get


/

-0
 2  3 -1 x
x
-- 1 x
/ -x

2
+2
2 + ln +
ln -B(
x) = 0
+
. (4.11)
2

1+x
/
3

1+x
/ For this illustrative example we choose the width of the slab so that B(
x) 0 if |
x| 1. This
corresponds to = 1.14. The coefficient is chosen to obtain a flat shape of the B(
x) curve given

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

65

/()

0.5

approx.
original

-0.5
-2

-1

x/b

Figure 4.7: Approximation of the original field distribution (Eq. (4.4)) with a field generated by the
current given by a 3-rd order polynomial.
by (4.11) over the range x
= [0.65; 0.65] (this range is chosen by inspection as the area between
two extremes typically presented on the B(
x) curve). The mathematical criteria of flatness is


0.65 
B(
x) Bav 2
d
x min
(4.12)
Bav
0.65
where
Bav

= 0
2 2 0.65

0.65
0.65

B(
x)d
x = 0

(0.84 + 1.77)
2

(4.13)

The integral (4.12) is minimized with 1.5. Substituting this value into (4.13), we get
Bav = 3.03

.
2

Since we want Bav = B =


=

(4.14)
B0
b x,

we need

B0 X
2
B =
3.030
0.4820 b

The resulting current distribution is


/
 3 0
x

B0 X
+ 1.5
.
x) =
J1 (
0.4820 b

(4.15)

(4.16)

The magnetic field produced by this current distribution along with the original field distribution
given by equation (4.4) are shown in Figure 4.7. The force per unit length in the Y direction which
would act on the slab if the current distribution were indeed given by (4.16) would be


B 2 x
J1 (
x)B(
x)bd
x= 0

Fcont =
0.4820


 

1  2


x

B2
x
4
x
3
x+2
x 3.32 0 x
+ 1.5 3 d
+ 1.5 3 d

0
1
1

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

66

/()

0.5

approx.
original

-0.5
-2

-1

x/b

Figure 4.8: Approximation of the original field distribution (Eq. (4.4)) with a field generated by the
current given by a 3-rd order polynomial and point currents.
Note that this equation underestimates the force because the current given by (4.16) does not completely compensate the field variation in the region [b; b] and creates additional fields outside of
the region. More accurate field matching would require higher order terms in the current model
built over whole length of the slab in the X direction. For simplicity here we assume that the slab
length in the X direction is limited to 2b (this allow us not to worry about field matching outside of
the region [b; b]). To account for boundary effects and produce a field exceeding B given by
(4.4) over the whole range [b; b] (more than enough) we introduce two point current I(
x )
and I(
x + ) at the ends of the slab, where (
x) is Diracs delta function. The magnetic field
produced by these two point currents in XY plane is




1
0 I
1
1
1
0
I
+
=
+
.
(4.17)
Bpoint =
2
b x b + x
2 b x
+x

We want this field to be equal to B at |x| = b (|


x| = 1). From this condition we find
I=

2 1
.
Bb
0

(4.18)

The equation (4.17) then becomes




1
2 1
1
Bpoint = B
+
2
x
+x

(4.19)

In order not to exceed the field value of B by too large an amount we reduce the coefficient .
An acceptable reduction factor was found to be 0.8. The resulting field produced by point currents
(4.18) and by the continuous current distribution (4.16) is shown in Figure 4.8. The force per unit
length due to the two point currents I is
Fpoint = 2B0 I = 2B02

B2
2 1
= 1.65 0 X
0
0

(4.20)

Summing up, the total force per unit length acting on the slab does not exceed
Fslab = 0.8Fcont + Fpoint 4.3

B02
X.
0

(4.21)

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

67

Figure 4.9: A conducting loop formed by two infinitely long round wires.
As an alternative to the slab, consider a conducting loop of infinite length in the Y direction as
in Figure 4.5 formed by two round wires (Figure 4.9). Both wires are located in the area of uniform
fields B0 . The wire cross section is chosen to simplify calculations.
The magnetic field within the loop is a superposition of the fields produced by each of the wires.
We assume that the current is distributed uniformly within the wire and that the magnetic flux
through the loop is the flux passing between the two wire centerlines. First, we find the magnetic
field generated on the X-axis by one of the wires. From the system symmetry it is clear that this
field is directed along the Z-axis. Considering the circulation of the magnetic field along a circle
with radius r = x, and using the Maxwell equation H = j, we find that
/
I
x if x < r0
2r02
H1z =
(4.22)
I
if x r0
2x
where I is the current through the wire and r0 is the wire radius.
Integrating (4.22) from 0 to l0 and multiplying the result by the factor of two to take into account
that there are two wires , we find the flux through the loop per unit loop length to be

r0




l0
1
l0
0
I
+ ln
H1z dx +
H1z dx =
= 20

2
r0
0
r0
Thus the inductance per the unit length of the loop is


0 1
l0

+ ln
L1 = =
I
2
r0

(4.23)

Then the current in the loop displaced by X is


I=

2B0 X
L1

(4.24)

The smaller L1 , the large the current. From (4.23), the smaller l0 is, the smaller L1 becomes. The
minimal feasible l0 is equal to 2r0 . To be conservative we take l0 = 3r0 . Then the minimal value
of L1 is
L1min = 1.6

(4.25)

The force acting on the loop in this case would be


Floop = 2IB0 = 7.9

B02
X
0

(4.26)

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

68

This is almost two times larger than the above rough upper estimate of the force in case of the slab
(equation (4.21)).
Note, that even the conducting slab example is rather advantageous as compared to the scheme
of a typical eddy-current bearing, which essentially consists of a magnet moving with respect to a
bulk conductor. Let us clarify this point. When a conductor interacts with a magnetic field, the force
F acting on it is given by the volume integral:

F =
j Bdv
(4.27)
To obtain high force magnitude at a particular point, we need high j, high B and we also require
that j be orthogonal to B. The last condition clearly does not have to be satisfied in the case of a
bulk conductor. However, it is satisfied in case of the conducting slab shown in Figure 4.6 and the
conducting loop (Figure 4.5).
Furthermore, to maximize F we need all the elementary forces j B be oriented in the same
direction. Again, this is the case for the conducting slab in Figure 4.6 and the conducting loop in
Figure 4.5, both of which are exposed to the field orthogonal to their planes. However, this is not
the case when a bulk conductor is used. The elementary force components which do not contribute
to the net force simply cause additional stresses in the conductor and waste its current-carrying
capacity (in principle we want to obtain the necessary force with as little current as possible, since
the current is associated with resistive losses).
The difference between the conducting-slab and conducting-loop examples is that in the latter,
B is higher because the currents are forced to flow where the fields are high.
Thus, the hierarchy is that the conducting loop system (Figure 4.5) is better than the conducting
slab system (Figure 4.6) and both are better than simply a bulk conductor interacting with a highgradient field.
A similar situation occurs with superconducting bearings: bearings utilizing interaction of a
bulk, hard, type II superconductor with a permanent magnet are structurally identical to an eddycurrent bearing that utilizes bulk conductors.
Another point of view on the problem is that j B is actually a force acting on the electrons
which can move freely within the conductor. When they experience this force, they simply move
in the direction of the force. The force is transmitted on the conductor only because the electron
motions are limited by the conductor boundaries. By careful design of these boundaries we keep
the electrons in the areas where the field and, consequently, the force are high.
Confirming the validity of these arguments, high load capacity and stiffness per unit surface area
(3.7 N/cm2 and 12 N/(mmcm2 )) have been obtained earlier in a prototype of a superconducting
magnetic bearing using superconducting loops instead of bulk superconductors [56].
In linear null-flux suspensions, this scheme has been proven to deliver much higher lift/drag
ratios ( 60/1) than if simple coils or aluminium slabs laid on the ground would interact with a
high-gradient magnetic field generated on the train (lift/drag 25/1) [75].
Another criterion for the choice of the rotating conductor shape is minimization of the angle ,
which leads to a lower minimal levitation speed. Some measures aimed to minimize , however, often cause undesirable decrease of the suspension stiffness. Thus, in case of a conductor G consisting
of separate loops, it can be shown (see Appendix B) that the angle is
= arctan

R
L

(4.28)

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

69

Figure 4.10: Geometric parameters of the conducting loops in Figure 4.4.


and the in-plane stiffness K is
(4BRav sin 0 )2

(4.29)
2L 1 + tan2
in which R and L are the resistance and inductance of
- each loop,
- n is the number of loops or order
z2 of rotational periodicity (three in Figure 4.4), B = - Bz1 B
, and the meanings of the geometric
2
parameters 0 and Rav are clear from Figure 4.10. It is easy to see that increase of the inductance
L reduces but also reduces the in-plane stiffness K.
Also, we have seen earlier that shaping the conductor as a loop, as in contrast to a slab, leads
to an increase of the suspension stiffness. On the other side, removing the central part of the slab
causes currents to flow longer distances through smaller cross-sections, which will likely raise the
apparent resistance and, consequently, may increase the angle .
The conducting loop shown in Figure 4.3 can be made as a shortened coil wound of several
turns of an insulated wire (tape). If the assumption of the uniform current distribution through the
wall cross-section of the original loop is true and the copper factor is unity, then this design change
will not affect the load capacity and stiffness of the suspension as well as the ratio between the loop
inductance and resistance, which determines the angle (see Appendix B).
To see this, let us assume that the cross-sectional area of the conducting wall of the original loop
was Sc , the average length of the loop was lav and its inductance and resistance were L0 and R0
respectively. Now we consider a loop made as a shortened coil comprising N turns of wire (tape)
filling the area Sc with the copper factor of . It is easy to see that the cross sectional area of the
wire in this case is Sc /N and its length from one end to the other is N lav . Then the resistance of
the multi-turn loop (measured between the wire ends) is
K =n

R = (1/)N 2 R0 .
Since the current distribution through the overall loop cross-section was and remains close to uniform, the average magnetic flux av , generated by the current I0 flowing through the entire cross
section of the original loop, is the same as the flux generated by the current I = I0 /N flowing
through the wire forming the multi-turn loop. The inductance of the original loop was L0 = av /I.
The inductance of the multi-turn loop is L = N 2 av /I. Thus,
L = N 2 L0 .
The ratio between R and L is then
1 R0
R
=
.
L
L0

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

70

When the copper factor is close to unity, this ratio remains constant under the described coil
modification.
The current in the original solid loop was the solution of the following differential equation:
L0

dext
dI0
+ R0 I0 =
,
dt
dt

(4.30)

where ext was the average external flux though the loop.
In case of the multi-turn loop, this equation becomes
L

dext
dI
+ RI = N
,
dt
dt

or
N 2 L0

1
dext
dI
+ N 2 R0 I = N
,
dt

dt

(4.31)

It is easy to see that the solution of this equation I can be presented as I = Ie /N , where Ie is
the current which would be observed in an equivalent solid conducting loop with inductance L0 and
resistance (1/)R0 .
Further we note that when a loop composed of N turns of wire carries a current I in each turn
and is exposed to a magnetic field B normal to the current, the Ampere force acting on the loop is
FA N BI = BIe .
Therefore, the electromagnetic force acting on a loop composed of N turns of insulated wire is the
same as the force acting on an equivalent solid loop with inductance L0 and resistance (1/)R0 . It
can be said that the copper coefficient increases the specific resistance of the loop = R0 /Sc by a
factor 1/, otherwise the multi-turn and solid loops are identical.
Note, however, that from the electromagnetic point of view, the multi-turn loop design has
several significant advantages. First of all, this measure will reduce eddy currents in the loop walls
when the rotor is displaced from radial equilibrium.
Second, in case of a multi-turn loop formed out of a fine wire the assumption of uniform current
distribution is guaranteed to be satisfied because of the current continuity along the wire. This may
not be the case, however, for a solid loop, especially at high frequencies when skin effects start to
0.5 for a copper conductor can be presented in
play an important role. (The
skin depth = [f ]
centimeters as Cu = 6.6/ f , where f is the field frequency, which in the proposed design is equal
to the rotational frequency. Thus at 100 Hz rotational frequency the skin depth is only 6.6 mm.)
Because of the skin-effect the resistance of a solid loop may turn out to be much bigger than the one
which could be expected considering the cross-sectional area.
Furthermore, in the case of a multi-turn loop a design opportunity appears to change favorably
the ratio of R/L and, subsequently, the angle . This can be achieved by connecting an inductor in
series with the windings as shown in Figure 4.11. If this inductor is located away from the magnetic
fields, it will not influence the static force characteristics of the suspension.
A possible method of manufacturing of a rotor with multi-turn loops using printed circuit board
technology is indicated in Figure 4.12.
To minimize the rotational loss, it is desirable to reduce the amount of the conducting material
in the area of the field gradient. If the rotating conductor is represented by isolated loops (coils)
with conducting walls of uniform thickness, then the number of loops should be minimal, which is

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

Figure 4.11: Multi-turn loop with additional inductance.

Figure 4.12: Forming a multi-turn loop using PCB technology.

71

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

72

Figure 4.13: Squirrel-cage rotor.


three. This measure, however, will negatively effect the angle , the suspension stiffness and the
rotor strength. Another variant of making the rotor (squirrel-cage rotor) is shown in Figure 4.13. In
this case, along with the optimization of the amount of conducting material in the radial sections
of the rotor it appears to be beneficial from the loss minimization point of view to make the radial
parts as narrow as possible. (Skin effect has to be kept in mind if this variant of the rotor is used.)
An interesting and practical observation is that the angle can be changed favorably by simply scaling up the system. As a simple example consider the inductance and resistance per unit
length of the loop shown in Figure 4.9. The inductance per unit length is given by (4.23) and the
corresponding resistance is
R1 = 2

1
r02

(4.32)

where is the conductivity of the wire material and the factor 2 appears because there are two wires
connected in series. Note that L1 does not depend on the actual values of r0 and l0 but only on their
ratio. If we scale the system proportionally, L1 will not change. In contrast, R1 will be reduced by
the second power of the scaling factor. Thus, by scaling the system up we can reduce the ratio R/L
and the angle correspondingly. Note that the angle would be zero if we used superconductors
with essentially infinite conductivity. Thus, the behavior of the rotating loops made of normal
conductors in this system approaches the superconducting limit when either rotational speed or
system size increases.
To calculate the rotational eddy-current loss occurring when the rotor is displaced radially, it
is useful to note that if the field can be approximated by the shape shown in Figure 4.5, then the
portion of the rotating conductor exposed to the high-gradient field will experience harmonic field
variations, the amplitude of which would be almost uniform along the radius. The methods developed earlier for calculating rotational loss in the laminations of the transformers and active magnetic
bearings [111, 112] can be adapted to estimate losses in this case.

4.3

Generation of the stationary magnetic field

In [55] and [56], the magnetic field was generated using the permanent magnet arrangement shown
in Figure 4.14. Figure 4.15 shows the distribution of the z-component of the magnetic field in the
x
z = Bz , x
middle of the air gap of this system, as calculated in the dimensionless units B
0 M = b

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

73

Figure 4.14: A variant of the magnetic system used to generate a magnetic field as shown in Figure 4.4.

Figure 4.15: Magnetic field distribution in the middle of the air gap of the magnetic system shown
in Figure 4.14
for g = 0.8b, = 0.5b, h = g2 . Here M is the magnetization of the magnets (0 M 1.2 T for
NdFeB).
The choice of the magnet thickness in this system has been made so as to maximize the ratio of
the second power of the magnetic field amplitude vs. magnet thickness in case when pole shoes are
installed on the magnet surfaces. Magnetic fields up to Bz =0 M 1.2 T can be obtained using
this simple arrangement, but when Bz approaches 0 M, magnet thickness h goes to infinity. The
Bz = 0 M limit can be exceeded if field concentration means are used, but it will inevitably cause
significant stray fluxes and likely will result in an inefficient usage of magnetic material.
It is easy to see that the magnetic field distribution has two extremum (a minimum and maximum) and, as expected, the field gradient is low in the proximity of these extreme points. These
areas of low-gradient fields simulate the areas of constant fields Bz = B0 shown in Figure 4.5.

4.4

Designing the stationary conductors and rotating magnetic field

In case of the stationary conductor G , neither its shape, nor the shape of the magnetic field with
which it interacts influence the rotational drag as long as the field is uniform circumferentially.
Structural properties of this conductor are also not so important since it does not rotate. The same

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

74

Figure 4.16: An electronic schematic to reduce the apparent resistance of a stationary coil.

Figure 4.17: Voltage distribution along the path of the current flowing through the stationary coil.
ideas as for rotating conductors and stationary fields can be applied to increase interaction forces.
As shown above, several parameters of the proposed suspension including minimal stable levitation speed depend on the resistance and inductance of the stationary loops. Note that these loops
are stationary and, therefore, easily accessible as in contrast to the rotating loops. Taking advantage
of this circumstance, a simple electrical/electronic circuit can be connected in series with the loops
to vary their resistances and/or inductances. For example, the circuit shown schematically in Figure 4.16 effectively reduces the loop resistance without influencing its inductance. The operation of
this circuit is more effective if a loop is wound of many turns of insulated wire. Because of this, we
will use the term coil rather than a loop when referring to this structure.
The path of the current flowing through the stationary coil is shown in Figure 4.17.
From the balance of the voltages and electromotive forces along this path, we find that the
voltage on the output of the operational amplifier (point b) is
Ub = I(Rcoil + R1 ) E

(4.33)

where I is the current in the loop and E is the electromotive force induced in accordance with
Faradays law whenever the rotor moves laterally: E = d
dt . Here is the net magnetic flux
through the loop, which includes the flux ext generated by the magnetic means G mounted on the

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

75

rotor and the flux LI induced by the current I. Thus,


E =

dext
dI
L
dt
dt

(4.34)

The output of the operational amplifier Ub is proportional to the difference between the voltages
applied to its two inputs with some coefficient k:


R3
Ub = K(Ua Uc ) = k IR1 Ub
R2 + R3
Thus


Ub 1 + k

R3
R2 + R3

= kIR1

If R3 = 0 and k ,
Ub

R3
= IR1
R2 + R3

(4.35)

Combining equations (4.33) and (4.35), we find the current through a stationary coil is
I=

E
R2
Rcoil R1 R
3

(4.36)

According to this equation, the current in the loop is the same as it would be if the loop had the
same inductance L but the resistance equal to the equivalent value
Reqv = Rcoil R1

R2
R3

(4.37)

Note that by choosing parameters R1 , R2 and R3 , one can make this equivalent resistance close
to zero or even negative. Resolving equation (4.36) with respect to E and substituting the final
expression into (4.34), we obtain the following equation for the current through the damping coil:
L

dext
dI
+ Reqv I =
dt
dt

(4.38)

Clearly, negative values of Reqv would make this system unstable. In practice, how small Reqv
can be made is limited by variations of other system parameters: under all operational conditions
the system must remain stable. Therefore, it is impossible in practice to make an exact analog
of superconducting bearings by making Reqv = 0. Note that if it happened, the damping force
would become a positioning force as in superconducting bearings. If Reqv = 0, the equation (4.38)
becomes
L

dI
dext
=
dt
dt

which after integration gives us


1
I = ext
L

(4.39)

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

76

where ext is a change of the magnetic flux that occurs when the rotor is shifted from the equilibrium position where I = 0. As known from studies of superconducting bearings, with Reqv = 0,
stability is possible without the second part of the system (rotating conductor G and stationary field
source Q). However, the equilibrium position in this case is not well-defined. Thus, in the case of
superconducting bearings, a rotor can be in equilibrium in any position, where the superconductors
were switched into the superconducting state (through cooling below the critical temperature). After this, physical effects such as temperature-activated flux creep may cause the rotor equilibrium
to drift with time. In case of hard type II superconductors much bigger ambiguity of the equilibrium position appears because of the remagnetization hysteresis. In contrast, in the proposed design
with Reqv > 0, the equilibrium is unique and well-defined: it occurs when the rotor rotation axis
coincides with the symmetry axis of the magnetic field generated by the stationary source Q. This
equilibrium is determined by the interaction between the conductor G and the magnetic field generated by the field source Q. The interaction between G and Q is needed to make the system stable.
Besides it can also effectively counteract dynamical radial loads provided that their magnitudes and
durations are such that the resulting radial displacements of the rotor do not exceed tolerable limits.
It is interesting to compare operation of this bearing and operation of an active magnetic bearing
from the energy point of view. Multiplying both sides of equation (4.39) by I, and integrating in
time we obtain

t
dI
dext
IL
dt0 =
I
dt0
(4.40)
dt0
dt0
0
0
t
ext
The right-hand term 0 I d
dt0 dt0 represents the work of the electromotive force induced during
the rotor motion.
external forces moving the
t ThisdI is equal with the negative of the work of the LI
2
rotor. The term 0 IL dt
dt
represent
variation
of
the
magnetic
energy
0
2 stored in the coil. Thus,
0
variation of the magnetic energy in this system is caused by the forces moving the rotor and not by
control systems. The electronic unit is only responsible for suppressing the coil resistance (if Reqv
were equal to zero, the work of the power supply feeding
t the stationary coil would be exactly equal
to the energy dissipated in the coil in form of heat: 0 I 2 Rcoil dt0 .) If the rotor axis does not move
radially for periods of time long compared to the time constant = L/Reqv , the electronics do no
work.
In spite of some active circuitry being involved, one can see that the operational principle of this
bearing is very different from the ones of conventional active magnetic bearings. The electronics
used here do not contribute to the restoring force, and they have a much smaller power rating. They
never increase either the mechanical or the electromagnetic energy in the system, but rather help to
dissipate the energy in the unwanted lateral motions of the rotor. Based on the arguments above, the
system appears to be passive even when electronics are introduced.
Another parameter of the stationary coils which can be varied using external circuitry is the
coil inductance. As shown in the previous chapter, the minimal levitation speed in the proposed
type of bearing can be reduced through an appropriate choice of the ratio of the stationary coil
inductance vs resistance. When air-core coils are used, this ratio typically needs to be increased.
A simple practical way of achieving this is to connect an additional inductor with a low-resistance
winding on a soft-magnetic core as in the case of the rotating coils (Figure 4.11). In situations where
additional electronic circuitry connected to stationary coils does not present a design problem, the
additional inductors can be simulated by electronic means.
An additional advantage of increasing the stationary coil inductance is that it leads to a decrease
in the rotational losses of an electromagnetic nature caused by the rotor unbalance, as shown in

CHAPTER 4. TECHNICAL AND PRACTICAL CONSIDERATIONS

77

Section 3.7.
As a next step after changing resistances and inductances, one can design electronic units connected to the stationary coils in a more sophisticated way as to reduce the minimum stable levitation
speed or achieve some other control goal. The advantages of this system (which could be active)
compared to a conventional active magnetic bearing would include fewer incoming wires (there is
no position sensor) and potentially lower power rating of the control circuits. Further, the active
components would be needed only at low speeds and might reasonably be switched out of operation
at high speed or under particularly quiescent operating conditions.
As a next possible step in relaxing the condition of minimal wiring, in addition to the stationary
coils considered above one can use separate monitoring coils which will produce a voltage proportional to the rotor speed. This voltage then will be used to control currents in the force-generating
coils. This scheme is more robust then the negative resistance scheme considered above. Of course,
one can also add an additional rotor position sensor.

Chapter 5

Experiments
To verify the theoretical results we have built and tested a prototype of this novel form of suspension
system. The major objectives of this experimental work were to prove the validity of the operational
principle, demonstrate the ability of the above presented theory to predict with a reasonable accuracy the minimal stable levitation speed and rotational losses. We did not pursue high levels of
suspension stiffness, high rotational speeds, or the minimization of rotational losses at this stage.
The rotational speed was limited mainly because of the rotor design, the strength of the rotor
material (it was made of phenolic) and an intent to use a simple low speed drive for the rotor.
The choice of both the design and the material of the rotor was dictated by the ease and cost of
manufacturing. The suspension stiffness was limited primarily because of the low rotational speed
and because of the ratings of the magnets used to generate the stationary magnetic field. (As such
we used inexpensive ceramic Grade 5 magnets with maximal energy product BHmax = 27 kJ/m3 ,
which all together cost about $20. More expensive N dF eB have BHmax = 320 kJ/m3 .)

5.1
5.1.1

Observation of stable non-contact levitation


Experimental arrangement

We have built a prototype of this novel form of suspension system, which is shown schematically in
Figure 5.1. In our prototype, by part number, we have four conducting multi-turn loops 1 connected
in series with additional inductors 2 and located at radius Rav =140 mm. The loops were constructed
of insulated flat ribbon wire having a 0.5x5 mm2 cross-section (supplied by MWS Wire Industries).
Each multi-turn loop includes 17 turns of wire. The added inductor includes a magnetic yoke
3 assembled of 0.35 mm thick transformer steel laminations with a total cross-sectional area of
5x9.5 mm2 and an air gap of about 0.2 mm. The inductor is wound from 36 turns of 0.5x10 mm2 flat
ribbon wire. The overall resistance and inductance of the assembly were measured to be 0.054 and
0.48 mH at 60 Hz. The loops are exposed to a magnetic field generated by ring permanent magnets
4 and 5. Steel disks 6 provide return paths for the magnetic flux. The measured distribution of the
axial component of the magnetic field is shown in Figure 5.2.
A photograph of the rotor is shown in Figure 5.3. The rotor outer diameter is 37 cm, the minimal
thickness on the outer boundary is 5 mm, while the maximal thickness is 17 mm (in the middle).
The rotor mass was measured to be 3.2 kg. A disk-shaped NdFeB magnet 7 of 76 mm diameter and
12.5 mm thickness mounted on the rotor generates the magnetic field that provides radial damping
when interacting with the six coils 8 mounted on the stator around its axis.
78

CHAPTER 5. EXPERIMENTS

79

Figure 5.1: Cross-sectional schematic view of the bearing prototype.

Figure 5.2: The magnetic field distribution as measured in the middle of the air gap of the prototype.

11
2
12

Figure 5.3: The rotor of the prototype. Part numbers shown are consistent with Figure 5.1.

CHAPTER 5. EXPERIMENTS

80

Cords restricting
radial motion of
the rotor

There is a disk-shaped
magnet 7 inside the rotor

Rotor

Stationary
lift magnet 9

Figure 5.4: The rotor being suspended vertically due to interaction between magnets 7 and 9. It is
unstable in the radial direction. The cords shown prevent it from slipping sideways but do not exert
significant force in the vertical direction.
We used the circuits shown in Figure 4.16 with the stator coils 8 to reduce their apparent resistance. With R1 =1 , R2 =12 k and R3 =1 k we reduced the coil resistance (which originally was
2
13 ) by R1 R
R3 =12 ). The resulting apparent coil resistance was 1 (13-times reduction).
Magnet 7 when repelling lift magnet 9 also provides axial suspension of the rotor. Lift magnet
9 is assembled from twenty-six 12.5 mm-thick rectangular NdFeB magnets with area perpendicular
to the magnetization direction of 12.5 mm x 6.3 mm. The rotor weight of 3.2 kg is compensated
when the distance between magnets 7 and 9 is equal to approximately 3 cm.
Figure 5.4 shows the rotor being suspended by means of magnets 7 and 9. The suspension is
clearly unstable in the radial direction and radial cords are used to prevent the rotor from slipping
in the radial direction. They do not exert much force in the axial direction and, in fact, even pull the
rotor down a little bit as can be seen in the photograph of Figure 5.4.
When the rotor is placed in the housing, its lateral motions are restricted by bushings 10 and
journals 11 until it reaches its minimum stable speed and moves to the central position.
Eight NdFeB magnets 12 with 6 mm diameter and 5 mm thickness provide some suspension
stiffness with respect to angular deflections about the radial axes when interacting with magnets
4. Due to the location of magnets 12 on the outer perimeter of the rotor, the restoring torque is
significant even though the interaction between magnets 12 and 4 is not particularly strong (0.3 N
maximum).
The bearing prototype has large air gaps, which allow the rotor be displaced by 2 mm in any direction, and it has relatively low stiffness, approximately 2 N/mm in the axial direction and 1 N/mm
in the radial direction at 30 Hz. For comparison, air gaps in active magnetic bearings are measured
in micrometers. Having large gaps is an advantage because it makes the device less sensitive to
intra-gap contamination. According to our theory, the stiffness values for the current size of the
bearing could be increased dramatically (up to 200 N/mm) if the ceramic magnets currently used in
the prototype were replaced with much more powerful NdFeB magnets and some other adjustments
were made. However, we did not pursue this goal at this stage; our intentions were to demonstrate
the possibility of non-contact suspension in this type of system and to verify the theoretical model.
Figure 5.5 shows a test rig used to carry out our experiments on passive levitation. In this rig,
the bearing rotor is accelerated using a variable speed motor connected to it through a flexible shaft,
which exerts almost no radial force on the rotor but is capable of delivering a driving torque. The
shaft is connected to the rotor through a threaded coupling. During acceleration, the torque applied
to the rotor tightens the connection. When the desired rotational speed is reached, simply switching

CHAPTER 5. EXPERIMENTS

81

Variable
speed DC
motor
Flexible shaft

Thread
coupling
Bearing
prototype

Figure 5.5: An overall view of the test rig.


off the motor power supply will reverse the torque (because now the motor lags and the rotor leads),
with subsequent disengagement of the coupling. This feature allowed us to investigate the bearing
behavior during free rotational run-down.

5.1.2

Results and discussion

Stable non-contact suspension of our 3.2-kg rotor was observed at rotational speeds above approximately 18 Hz. This critical speed value was measured during the rotor run-down. After the rotor
slowed below this critical speed, it started a spiral motion from the central position until the shaft
11 touched the back-up bearings 10. Approximately the same critical speed value was measured on
run-up but it was harder to detect precisely.

5.2

Measurement of the radial suspension stiffness

We have measured the radial suspension stiffness developing in the bearing when its rotor rotates
above the minimal stable levitation speed.
To obtain radial loading on the rotor of the bearing prototype, we tilted the prototype about one
of the horizontal axes (Y-axis) by a certain angle. This caused a projection of the gravity force on
the perpendicular X-axis. The tilting was achieved by lifting one point of the housing of the bearing
by means of a z-table, while hinging the diametrically opposite point. The distance between the
two points was 391 mm. With the rotor mass 3.2 kg, the radial force acting on the rotor in the X
direction as a function of z was
F = 0.0802z,
where F is in Newtons and z is in millimeters.

(5.1)

CHAPTER 5. EXPERIMENTS

82

Figure 5.6: Radial equilibrium force diagram.


The radial displacements of the rotor in two orthogonal directions were measured with two
optical displacement sensors (see Appendix E). The rotational speed was monitored with a handheld optical tachometer.
Figure 5.6 shows the radial equilibrium force diagram when the rotor is subjected to an external
radial force Fext acting in the x-direction. The force component Kdes r appears because of the
destabilizing influence of the axial suspension in the radial plane.
It is easy to see that the angle between the X axis and the rotor displacement is equal to (by
definition, is the angle between the direction opposite to the rotor displacement and the electromagnetic force caused by this displacement); the proportionality coefficient between the external
force and the rotor displacement is equal to K ; and the parameters of the radial suspension without
the axial suspension, K and , can be calculated as

2 + 2K K

K = K 2 + Kdes
(5.2)
des cos ,
and
= arctan

K sin
.
+ Kdes

K cos

(5.3)

The results of the measurement of the radial destabilizing stiffness caused by the axial suspension, Kdes , are shown in Figure 5.7. The experimentally measured value of Kdes was found to be
equal 1.022 N/mm.
The radial force displacement characteristics of the bearing measured at different speeds are
provided in the Appendix F.
Some predictions of the radial suspension parameters K and were made before building the
prototype based on the following equations:
K=

n (4N BRav sin 0 )2

,
2 L 1 + tan2

(5.4)

CHAPTER 5. EXPERIMENTS

83

3
Experiment
Fit

2.5

F, N

2
1.5
1
0.5
0

0.5

1.5

X, mm

2.5

Figure 5.7: Results of the measurements of the radial destabilizing stiffness Kdes introduced by the
axial suspension. The fitting function is F = 1.022x.

= arctan

R
L


.

(5.5)

(The first equation is an extension of the equation (B.10) to the case when the loop comprises N
turns of the wire (tape); the second equation is the equation (B.9) repeated here for convenience.)
The Table 5.1 contains the information about the prototype, which is necessary for calculating
K and .
The predicted and measured values of K and are in a reasonably good agreement, as shown
in Figures 5.8 and 5.9.
The difference between the predicted and measured values of K does not exceed 15% of the
predicted value. The difference between the predicted and measured values of the angle does not
exceed 30% of the predicted value.

5.3

Measurement of the rotational losses

We extracted the rotational losses in the prototype from the run-down rotational speed measurements. During run-down in air, the rotor speed was observed to decrease from 30.4 Hz to the
critical value of 17.7 Hz in 83 seconds . With the rotor moment of inertia being 0.031 kgm2 , the
average rotational energy loss rate in this frequency range was estimated to be 4.5 W. We believed
that most of the energy loss was due to air resistance and we repeated the experiment in a rough
vacuum, at a pressure of approximately 0.75 mm of mercury. The speed was observed to fall from
45.05 Hz to 18.86 Hz in 450 seconds, suggesting an average loss rate of 2.3 W.
Three sources of the rotational loss were expected to contribute significantly to the overall loss
value: eddy current couplings induced in the conducting housing and pole shoes that were exposed to the rotating, circumferentially non-uniform magnetic field emanating from magnets 12
(Figure 5.1); the rotor unbalance (misalignments between the axes which were assumed to be coincident in our analysis, see Section 3.7); and circumferential non-uniformities of the magnetic fields.

CHAPTER 5. EXPERIMENTS

84

Parameter
Number of coils, n
Number of turns in a coil, N
Average radius, Rav
Coil angle, 0
Magnetic flux density*, B
Measured inductance of a coil**, L
Measured resistance of a coil**, R

Value
4
17
140 mm
34.50 0.6 rad
0.13 T
0.481 mH
0.0541

Table 5.1: Parameters of the bearing prototype used for the prediction of its radial loading characteristics.
* This is the average absolute value of the magnetic flux densities in the areas of the conducting
arcs. Please refer to Figure 5.2 for the details of the radial field profile.
** L and R here represent the average values measured in four coils used in the prototype (see
Table E.1). Please refer to the Appendix E for the measurement details.

K, N/mm

1.6

1.2

0.8

Measured
Predicted

0.4

10

20

30

40

Rotational frequency, Hz
Figure 5.8: Predicted and measured radial suspension stiffness, K.

CHAPTER 5. EXPERIMENTS

85

40

30

20

Measured
Predicted
10

0
15

20

25

30

35

40

45

Rotational frequency, Hz
Figure 5.9: Predicted and measured angle .
The first source was simply a design drawback of the prototype (it was easier to use separate
magnets than continuous). To make this field more uniform we installed two pole shoes in the form
of steel rings on opposite sides of the rotor above the magnet poles. This measure, however, had
additional effects on the system.
Some effects were easy to measure or estimate. Thus, the addition of the steel rims increased
the rotor moment of inertia to 0.041 kgm2 and caused the negative radial destabilizing stiffness to
increase (in absolute value) to 1.25 N/mm from original 1 N/mm. The latter resulted in the
critical frequency increasing to 32 Hz.
The other effects were rather difficult to quantify. In particular, the addition of the steel rims
could change the radial equilibrium of the axial suspension system, i.e. the position of the rotor
where the axial suspension produces no radial force. (More specifically, it could change the location
of the axis Zar , see Section 3.7). The misalignment between the axes Zar and Z0 and between Zas
and Z0 likely causes additional rotational losses, but we did not carry out their theoretical estimate.
This can be suggested as a subject for future research.
When we repeated the experiment with the steel rims, the rotational speed fell from 48.01 Hz to
32.30 Hz in 570 seconds, suggesting an average loss rate of 1.4 W. Thus, addition of the rims was
definitely beneficial.
The rotor speed decays with time during run-downs in vacuum are shown in Figure 5.10. The
lower graph corresponds to the original system (without ring-shaped pole shoes linking the magnets
12 together) and the upper graph corresponds to the system with magnets 12 linked by the pole
shoes. The power losses calculated for both cases using the data shown in Figure 5.10 as functions
of rotational speed are shown in Figure 5.11. (We calculated the power losses using equation P =
2
J d
dt ).
In the available speed range, they appeared to increase almost linearly with the rotational speed.
The linear fitting equations however had negative intersects, suggesting negative power loss at zero
speed. To obtain zero power at zero speed we tried higher order fitting functions: pure quadratic
and a combination of linear and quadratic functions. The regression analysis was carried out using
the MathCAD built-in function linfit. The results of the approximations are shown in Figure 5.12

CHAPTER 5. EXPERIMENTS

86

Figure 5.10: Rotor speed decay during run-downs in vacuum.

Rotational power loss, W

Separate magnets 12
Magnets 12 linked by pole shoes

0
0

10
20
30
40
Rotational frequency, Hz

50

Figure 5.11: Rotational power losses during run-downs calculated using the data shown in Figure 5.10.

CHAPTER 5. EXPERIMENTS

87
Experiment
-0.791 + 0.098 f
2
0.0021 f
2
0.046 f + 0.0008 f

Rotational power loss, W

0
0

10
20
30
40
Rotational frequency, Hz

50

Figure 5.12: Approximations of the measured rotational loss in the experiment with separate magnets 12.
(for the experiment without the additional steel rims) and in Figure 5.13 (for the experiment with
the additional steel rims).
In the experiment without the additional rims, the loss was better approximated by a combination of linear and quadratic functions: Ploss = 0.046f + 8.062 104 f 2 . A pure quadratic function
could not give a close fit (the best fit in this case was obtained with Ploss = 2.078 103 f 2 ).
Interestingly, when the rims were added, the quadratic function (Ploss = 8.47 104 ) gave a
reasonably good fit. The combination function in this case was Ploss = 5.51103 f +7.135104 f 2 .
Thus, the addition of the rims reduced the linear loss component by almost an order of magnitude
while left the quadratic component almost intact. If we assume that eddy current losses vary as a
second power of the frequency, than addition of the rims had more influenced some other mechanism
of loss. As a hypothesis, it can be suggested that adding the rims had reduced the misalignment
between the axes Zar and Z0 .
If the remaining rotational losses are caused by eddy currents in the conducting parts of the
bearing, they will come to saturation with further increase of the rotational speed when the inductive
component of impedance becomes dominant. Replacement of the bulk aluminum housing with a
non-conducting one may result in further reduction of the rotational loss.
The design of the prototype allows easy estimation of the rotational losses caused by the currents in the stationary coils, which could be measured. Thus, if we neglect influence of the axial
suspension, this loss would be caused by the misalignment of the rotor mass center and the axis of
the magnetic field generated on the rotor Z0 . We estimated this loss in Section 3.7. The rotor was
balanced on a needle to make sure that the above misalignment was less than 1 mm, however its
exact value was not determined.
To estimate the loss using measured values of the currents in the stationary coils we multiply
both sides of equation (4.38) by I and integrate over the period of rotation T . Then we get

L
0

dI
I dt + Reqv
dt

I dt =

I
0

dext
dt
dt

(5.6)

CHAPTER 5. EXPERIMENTS

Rotational power loss, W

88
Experiment
-1.115 + 0.062 f
2
0.00085 f
2
0.0055 f + 0.0007 f

1.5

0.5

0
0

10
20
30
40
Rotational frequency, Hz

50

Figure 5.13: Approximations of the measured rotational loss in the experiment with magnets linked
by pole shoes 12.
T
0
When the current I is a harmonic function, L 0 I dI
dt dt is zero because of the 90 phase shift
between I and dI/dt. This integral represents the variation of the inductive energy, which is conserved. The right-hand term of equation (5.6) represents the work of the external electromotive
forces and is equal to the negative of the work performed by a moving magnet (or rotational energy
loss over the period T ). When divided by T this integral gives an average power loss over one rotor
revolution due to one stationary coil. To find the overall loss Ploss , we have to multiply this by the
number of coils N . Therefore,

1
1 T 2
Ploss = N Reqv
I dt = N Reqv I02 ,
(5.7)
T 0
2
where I0 is the amplitude of I.
In the run-down experiments, it was observed that I0 was approximately 0.07 A, essentially
independent of the rotational speed in the range 20 to 30 Hz (this measurement was conducted in the
air and higher speeds were hard to achieve because of the air resistance). Using equation (5.7), we
estimate the rotational loss due to the rotor unbalance as 0.015 W (we had 6 coils and Reqv 1 ).
This is really a negligible value.
Some theoretical estimates of this power loss, neglecting influence of the axial suspension, can
be obtained using the methods of the Section 3.7. If we neglect the stationary coil inductances, then
the rotational loss can be estimated using equation (3.139). In our prototype we had C 143 Ns
m
and m = 3.2 kg. Using equation 3.139 yields Ploss 2 W at 20 Hz assuming the unbalance of
1 mm. This is much higher than the measured value.
If we take the inductances of the stationary coils into account, then with the unbalance of 1 mm
using equation (3.149) we get Ploss = 0.23 W . Therefore, in the current prototype stationary coil
inductances cause a reduction of the rotational loss by almost an order of magnitude at 20 Hz rotational frequency. The difference becomes much bigger with the increase of the rotational frequency.
For example at 50 Hz the non-inductance estimate (3.139) would be 13.8 W, while the estimate
taking the inductance into account (3.149) would be just 0.43 W. This is already very close to the

CHAPTER 5. EXPERIMENTS

89

saturation loss value given by (3.151): 0.5 W. To match exactly the theoretical estimate of the rotational loss and its measured value at 20 Hz (0.015 W), we would need to assume that the unbalance
was 0.2 mm. This value does not appear to be too unrealistic, however it is difficult to expect a very
accurate match here mainly because we neglected the influence of the axial suspension and also
because of significant uncertainties in estimating C.
In general, in spite of low manufacturing accuracy and very rough balancing (1 mm) the
operation of the device was very smooth and noiseless. We observed no vibration after the motor
was disconnected. The rotational losses are reasonably low. This confirms the theoretical conclusion
that this type of bearing is very robust to manufacturing and balancing inaccuracies.

5.4

Effects of the stationary coil inductance on the


suspension properties

5.4.1

Prediction of the influence of the inductance on the minimum stable levitation


speed in the existing prototype

We attempted to show experimentally that the suspensions minimum stable levitation speed can be
reduced through an appropriate choice of the stationary coil inductance (more specifically, through
increase of the stationary coil inductances in the prototype). According to the data estimates used
in the prototype design, the angle becomes equal to /2 at a rotational frequency 16.5 Hz. The
resistance of the stationary coils was measured to be 13.55 0.55 and the inductance to be
17.00 0.75 mH. The electronic circuits connected to the stationary coils reduced their apparent
resistances to 1 , but did not influence the inductance. The resulting damping in the system was
estimated to be 143 Ns
m . Neglecting the inductance of the stationary coils, we expected the
minimum stable speed of the suspension to be 20 Hz. The time constant of each coil with the
electronic circuit connected to it was estimated to be 17 ms.
Figure 5.14 shows calculated curves indicating how much damping is needed for stability in this
prototype for various . These curves are obtained by presenting the curves shown in Figure 3.12
in a different set of variables: and C instead of and s.
As noted earlier, these curves are only approximations to the stability bounds obtained neglecting the speed cross-coupling terms in the system (3.89). The exact stability bounds calculated using
the algorithm presented in the Appendix A in most cases were hardly distinguishable from the approximate bounds. The biggest discrepancy was observed with = 0.06 (see Figure 5.15).
If we assume that our estimates of the prototype parameters are accurate, we may expect a
dramatic reduction of the critical speed from 18 Hz to approximately 11 Hz when increasing
the inductance of the stationary coils by 40 mH (from 17 mH to 57 mH) as can be seen in Figure 5.14
where a dashed line is drawn at an estimated level of C = 143 Ns
m . However, this reduction is very
sensitive to bearing parameter variations. For example, if the real damping coefficient happens to
Ns
be 120 Ns
m instead of 143 m , the bearing would be unstable with = 0.06 s (see Figure 5.14).
To make the situation even more complicated, we found that the minimum stable speed of our
bearing prototype varied in a wide range (10 Hz-20 Hz) after every disassembly/assembly cycle.
It was consistent, however, if we did not perform the disassembly/assembly procedure except that
the speed slowly increased if a series of tests was carried out continuously. This slow increase
in minimum stable speed can most likely be attributed to heating of the rotating conductors and a
corresponding increase in their resistances.

CHAPTER 5. EXPERIMENTS

90

Damping coefficient, Ns/m

500

=0.1
400

=0.06

300

=0

200
100
0

=0.04
10

=0.02

=/2
15

20

Rotational frequency, Hz

Figure 5.14: Stability regions for the bearing prototype

5.4.2

Experimental arrangement

To vary the inductance of the stationary coils in our experiment, we used commercially available
encapsulated toroidal inductors, model CMT908-H4 manufactured by MagneTek. These inductors
include two windings on a common toroidal core. The inductance and resistance of each winding
were measured (by a multimeter at low currents) to be 34.5 3.5 mH and 0.16 respectively. We
connected these windings in parallel with each other and in series with the stationary coils of the
bearing. The connection was made through a switch, the commutation of which allowed us to have
the inductor windings connected either in phase or in counter phase (see Figure 5.16). When the
windings were connected in phase, the inductance of the assembly was in excess of 31 mH (subject
to the fluctuations between different inductors from a batch). When the windings were connected
in counter phase, the inductance of the assembly became essentially zero. The assembly resistance
remained 0.08 in either case. This arrangement allowed us to quickly and conveniently change
the coil inductances.
The rotational speed of the rotor was measured using a signal from an optical pick-up monitoring
the angular position of the rotor. We also monitored the voltage on one of the stationary coils with
a digital oscilloscope which had a built-in FFT function. When the rotor was loosing stability on
run-down, the spectrum of the stationary coil voltage abruptly developed a large subsynchronous
peak which was very easily identified. The oscillograms and spectrograms of the measured signals
at the beginning of the transition and after it was complete are shown in Figures 5.17 and 5.18. For
comparison, Figure 5.19 shows the oscillograms obtained during the stable levitation of the rotor. It
is easy to see that in this case all the major peaks occur at multipliers of the rotational frequency.
The rotational frequency change during the transition was observed to be less than 0.3 Hz regardless of whether the additional inductors had been connected or not. The rate of the rotor slow
down right after the transition did not exceed 0.1 Hz/s. Assuming that the time between the moment
when the rotor hit the back-up bearing and the time when the operator picked up the rotational speed
was less than one second, the error due to the operator reaction time did not exceed 0.1 Hz. In total,
the upper bound on uncertainty in measuring differences between minimum stable levitation speeds

CHAPTER 5. EXPERIMENTS

91

Damping coefficient, Ns/m

600
500
400

Approximate
Exact

300
200
100
0

10

15

20

25

30

Rotational frequency, Hz
Figure 5.15: Comparison of the exact and approximate stability bounds calculated for = 0.06.

Figure 5.16: Two variants of the connection of the winding of the additional inductor.
Variant a.) non-zero inductance
Variant b.) zero inductance
Resistance remains the same in both cases.

in two experiment runs was less than 0.4 Hz.

5.4.3

Results and discussion

When we connected the additional inductors to all six stationary coils in the bearing prototype, we
found that the rotor center never comes to a stable position but instead moves in a circle around
the bearing center. There was still no mechanical contact with the back-up bearing when the rotor
rotated above a specific speed. As known from control theory [113], such a phenomenon (commonly
referred to as a limit cycle) is impossible in linear systems and is a manifestation of some non-linear
effects, which have not been considered in the simple analysis presented above. The reduction of
the additional inductance by connecting two inductors in parallel did not eliminate the limit cycle
completely. We did not pursue further reduction of the inductance since the effect of the inductance
would become quite difficult to detect.
Instead we tried to use another arrangement where additional inductors of 32 mH were connected only to three of the stationary coils (either upper or lower). We found that, with this arrange-

Amplitude, dB

Volts

Volts

CHAPTER 5. EXPERIMENTS

92

2
0
-2

0.5

1.5

0.5

time, sec 1

1.5

4
2
0

Rotational frequency

0
-25
-50 0

10

20

freq., Hz

30

40

50

Amplitude, dB

Volts

Volts

Figure 5.17: Oscillograms measured at the beginning of the transition from stable to unstable suspension.
The upper oscillogram - voltage across a stationary coil.
The middle oscillogram - signal from the optical pick-up used to monitor the rotational speed.
The lower graph - FFT of the upper oscillogram.
2
0
-2
-4

0.5

1.5

0.5

time, sec 1

1.5

4
2
0

Rotational frequency

0
-25
-50 0

10

20

freq., Hz

30

40

50

Figure 5.18: Oscillograms measured right after the transition from stable to unstable suspension.
The upper oscillogram - voltage across a stationary coil.
The middle oscillogram - signal from the optical pick-up used to monitor the rotational speed.
The lower graph - FFT of the upper oscillogram.

ment, the rotor gains a unique equilibrium position. Even though this case, strictly speaking, was
not considered in the above analysis, it can be used to demonstrate that the critical speed of the suspension can be reduced through increase of the stationary coil inductances. When we used another
set of inductors from the same batch with slightly larger inductances ( 38 mH) and connected them
to three coils, the limit cycle reappeared. The oscillograms measured in this case are shown in Figure 5.20. Note an additional peak in the spectrum below the rotational frequency, which normally
does not appear. Even with the rotor moving in a limit cycle, the increase of the coil inductances
seemed to result in reducing the minimum stable levitation speed, but the effect was smaller than
when inductances of only three coils were increased without producing a limit cycle.
We conducted twenty pairs of experiments, each including one run with additional inductances
on (connected to three of the coils) and one with additional inductances off. In order to eliminate the
effect of the bearing parameter creep with number of runs, we alternated the order of the inductanceon and inductance-off runs. In fifteen pairs the end of the transition to unstable motion was detected

CHAPTER 5. EXPERIMENTS

93

Volts

1
0
-1

Amplitude, dB

-2

0.5

Time, sec

1.5

Rotational frequency
0
-20
-40
-60

10

20

freq., Hz

30

40

Figure 5.19: Oscillograms measured during stable levitation.


The upper oscillogram - voltage across a stationary coil.
The lower graph - FFT of the upper oscillogram.

by the sound of the rotor hitting the back-up bearing. In five other pairs, we defined the minimum
stable levitation speed as the speed at the moment when a low frequency peak, that developed in
the spectrum of the voltage across one of the stationary coils, exceeded a 20 dB benchmark. The
spectra of one such pair are shown in Figure 5.21.
In all twenty pairs, the critical speed was detected to be lower with additional inductance. The
rotational speed difference was observed to fall in the range 0.87 0.29 Hz. The lower bound on
this difference is still higher than the above rough estimate of the upper bound on the experimental
uncertainty (0.4 Hz).
We noted that connection of additional inductors had significantly reduced the amplitude of the
spin-synchronous sinusoidal voltage on the stationary coil: the reduction was nearly a factor of two.
This voltage was most likely caused by rotor unbalance (more accurately by non-coincidence of
the rotor mass center and the symmetry axis of the rotor magnetic field). Thus, increase of the coil
inductance has an additional advantage of reducing power loss caused by unbalance.

CHAPTER 5. EXPERIMENTS

94

Volts

1
0

Amplitude, dB

-1

0.5

Time, sec

1.5

Rotational frequency

-20
-40
-60

10

20

freq., Hz

30

40

Amplitude, dB

Figure 5.20: Oscillograms measured when the bearing operated in a limit cycle.
The upper oscillogram - voltage across a stationary coil.
The middle oscillogram - signal from the optical pick-up used to monitor the rotational speed.
The lower graph - FFT of the upper oscillogram.

Without inductance
With inductance

0
-20
-40
-60

10

20

freq., Hz

30

40

Figure 5.21: Spectra of the voltage across one of the stationary coils measured in two consecutive
experiments: with and without additional inductance.

Chapter 6

Conclusions and Suggestions for Future


Research
This dissertation proposes a novel method of non-contact electromagnetic suspension and a novel
class of magnetic bearing, which operation is based on the above method.
The proposed method of non-contact suspension makes use of three types of interactions: dynamic interaction of a rotating conductor with a stationary circumferentially uniform magnetic field;
dynamic interaction of a rotating source of a circumferentially uniform magnetic field with a stationary conductor, and static interaction between two sources of time invariant magnetic fields.
The proposed type of magnetic bearing consist of two parts, rotating and stationary, each including sources of circumferentially uniform magnetic fields and circumferentially periodic conductors.
When there is no radial loading and the proposed bearing is in equilibrium, no change of the
magnetic field occurs in any point of the space (including conducting volumes) during the rotor
rotation, and, consequently, no electrical fields, E, and no currents develop. This leads to potentially
very low rotational losses. Because of this feature we named the proposed bearing a Null-E
Bearing. (The name is chosen by analogy with Null-Flux Bearings where no change of the flux
through conducting coils occurs at equilibrium in the absence of external loading.)
A general analysis of the interaction in this type of system, presented in the dissertation, consists
of the following steps and leads to the following conclusions:
1. Analysis of the radial electromagnetic forces acting on an arbitrary conducting body G exposed to a magnetic field Bz (r) generated by sources Q. The field is circumferentially uniform about some axis Z0 and is directed along this axis. The conductor rotates about some
axis Z parallel to Z0 . The sources Q are stationary.
The equations for the forces acting on the body G are derived.
2. Analysis of the dynamics of slow radial motions of body G superimposed on high speed
rotation about axis Z.
It is shown that if the slow motion assumptions are satisfied, the equilibrium that occurs when
axis Z coincides with axis Z0 is always unstable.
3. Analysis of the inverse system, including magnetic sources Q rotating about some axis Z0
and interacting with a stationary conductor G . The field generated by the sources Q is
95

CHAPTER 6. CONCLUSIONS

96

circumferentially uniform about axis Z0 .


It is rather surprising, but this system behaves completely differently from the system considered in step 1, even though they are identical structurally. The source of the difference is
clarified. The equations for the forces acting on Q are obtained assuming that the inductive
properties of the body G can be neglected.
4. Analysis of the radial motion of a combined body Q + G comprising bodies Q and G
arranged so that axes Z0 and Z coincide, interacting with a combined body Q+G comprising
bodies Q and G arranged so that their axes Z0 and Z  coincide. This analysis does not
consider the axial suspension, which can be provided using the static interaction between the
magnetic field sources Q and Q (as in Figure 1.2), or between separate sets of magnets.
This time it is shown that there is an in-plane equilibrium position for which axis Z (Z0 )
coincides with axis Z  (Z0 ) and this equilibrium can be made stable. The stability condition
is derived.
5. An analysis including axial suspension.
Here, a stability condition for the whole assembly is derived.
6. Modified stability analysis including inductive component of the impedance of the conductor
G . The analysis is carried out for a practically important case when the conductor G is
formed by a set of individual current loops with negligible mutual inductance.
It is shown that the inductive component may have a stabilizing effect and that the minimal stable levitation speed can be reduced through appropriate choice of the stationary coil
inductance.
Considering that the stationary conductors G are easily accessible, their currents can be controlled using external electronic circuitry, if it does not cause design or operational problems. In
contrast to the electronics of active magnetic bearings, these electronics do not necessarily have to
contribute to the restoring force and, therefore, may have a much smaller power rating. In particular,
these electronics can be used to suppress the resistive component and/or to increase the inductive
component of the stationary conductor impedances in order to reduce the minimal stable levitation
speed.
There are two major possible sources of the rotational loss in the proposed bearing: field circumferential non-uniformity and non-coincidences of the system critical axes (unbalance). The
equations to calculate the contribution from the second source are derived in the dissertation neglecting influence of the axial suspension.
To verify theoretical results, a suspension system has been built and tested, in which a 3.2-kg
rotor is suspended without mechanical contact when it rotates above approximately 18 Hz. The
major conclusions of the theoretical analysis have been corroborated by the measured behavior of
this apparatus.
The proposed class of magnetic bearing has a significant potential for applications because it
features a unique combination of properties which are not collectively provided by any of the prior
technologies. These include:

CHAPTER 6. CONCLUSIONS

97

1. Very high reliability due to intrinsic stability;


2. Wide temperature operation range including room temperature;
3. High efficiency;
4. Load capacity and stiffness sufficient for many applications;
5. Low rotational loss, virtually zero if only axial loading is applied - a condition which is easy
to satisfy in stationary applications;
6. No connecting wires and power supplies, unless supplementary electronics are used to augment system performance;
7. If supplementary electronics are used, their power rating will be much lower than in active
magnetic bearings;
8. Low price;
9. Low requirements for manufacturing accuracy.
Possible applications of the proposed magnetic bearing include flywheel energy storage systems
and turbomolecular pumps.
For the future research, the following can be suggested:
1. Analysis of the system dynamics without making a high-rotational speed assumption.
2. Analysis of the manufacturing inaccuracies on system stability.
3. Further development of the loss model.
4. Experimental demonstration of stable levitation without any electronics.
5. Development of control algorithm to control currents in the stationary coils in order to achieve
the specific control objectives.
6. Analysis of problems relating to the integration of this type of bearing into flywheels, turbomolecular pumps and other devices where it can be used.

Bibliography
[1] Malcolm McCaid, Permanent Magnets in Theory and Practice, Pentech press, London:
Plymouth, 1977.
[2] Peter Campbell, Permanent Magnet Materials and their Applications, Cambridge University
Press, New York, 1994.
[3] Jean-Paul Yonnet, Guy Lemarquand, Sophie Hemmerlin, and Elisabeth Olivier-Rulliere,
Stacked structures of passive magnetic bearings, Journal of Applied Physics, vol. 70, no.
10, pp. 66336635, 1991.
[4] S. Earnshaw, On the nature of the molecular forces, which regulate the constitution of the
luminiferous ether, Transactions of Cambridge Philosophical Society, vol. 7, pp. 97112,
1842.
[5] Richard Haberman, Elementary Applied Partial Differential Equations, Prentice Hall, Upper
Saddle River, 1998.
[6] W. Braunbek, Freischwebende korper im elektrischen und magnetischen feld, Z. Phys.,
vol. 112, pp. 753763, 1939.
[7] W. Braunbek, Freies schweben diamagnetischer korper im magnetfeld, Z. Phys., vol. 112,
pp. 764769, 1939.
[8] John S. Bay, Linear State Space Systems, WCB/McGraw-Hill, 1999.
[9] F. Matsumura, Y. Okada, M. Fujita, and T. Namerikawa, State of the art of magnetic bearings, JSME International Journal, vol. 40, no. 4, pp. 553560, 1997.
[10] J. B. Breazeale, C. G. McIlwraith, and E. N. Dacus, Factors limiting a magnetic suspension
system, Journal of Applied Physics, vol. 29, pp. 414, 1958.
[11] A.G. Lautzenhiser, Three-axis magnetic suspension, US Patent 3 146 038, Aug. 25, 1964.
[12] P.J. Gilinson and M. Chelmsford,
Patent 3 184 271, May 18, 1965.

Supports for rotating or oscillating members, US

[13] J. Lyman, Virtually zero powered magnetic suspension, US Patent 3 860 300, Jan. 14,
1975.
[14] H Ishikawa, M. Sei, and Eguchi I., Magnetic bearing devices, US Patent 3 909 082,
Sept. 30, 1975.
98

BIBLIOGRAPHY

99

[15] E.H. Maslen, C.K. Sortore, G.T. Gillies, R.D. Williams, S. J. Fedigan, and R. J. Aimone,
Fault tolerant magnetic bearings, ASME Journal of Engineering for Gas Turbines and
Power, vol. 121, no. 3, pp. 504508, 1999.
[16] M.D. Noh and E.H. Maslen, Self sensing active magnetic bearings based on parameter
estimation, IEEE Trans. Instrum. Meas., vol. 46, no. 1, pp. 4550, 1997.
[17] E. H. Maslen, P. E. Allaire, M. Noh, and C. K. Sortore, Magnetic bearing design for reduced
power consumption, ASME Journal of Tribology, vol. 118, no. 4, pp. 839846, 1996.
[18] E. H. Maslen, P. Hermann, M. A. Scott, and R. R. Humphris, Practical limits to the performance of magnetic bearings: Peak force, slew rate, and displacement sensitivity, ASME
Journal of Tribology, vol. 111, no. 2, pp. 331336, 1989.
[19] David W. Lewis, Robert R. Humphris, Eric H. Maslen, Paul E. Allaire, and Ronald D.
Williams, Magnetic bearings for pumps, compressors and other rotating machinery, US
Patent 5 355 042, Nov. 13, 1990.
[20] David W. Lewis, Robert R. Humphris, Eric H. Maslen, Paul E. Allaire, Ronald D. Williams,
and Steven W. Yates, Magnetic bearing systems, US Patent 5 347 190, Sept. 13, 1994.
[21] M. B. Scudiere, R. A. Willems, and G. T. Gillies, Digital controller for a magnetic suspension system, Rev. Sci. Instrum., vol. 57, no. 8, pp. 16161626, 1986.
[22] S. Sasaki, Stabilization of precession-free rotors supported by magnets, Journal of Applied
Physics, vol. 62, no. 7, pp. 26102615, 1987.
[23] B. E. Bernard and R. C. Ritter, The frequency fluctuations of an electronically cooled rotor,
Journal of Applied Physics, vol. 64, no. 6, pp. 28332837, 1988.
[24] M. A. Lawson and G. T. Gillies, Interrupt driven digital controller for a magnetic suspension
system, Rev. Sci. Instrum., vol. 60, no. 3, pp. 456466, 1989.
[25] T. Ohji, S. C. Mukhopadhyay, M. Iwahara, and S. Yamada, Permanent magnet bearings for
horizontal and vertical shaft machines: A comparative study, Journal of Applied Physics,
vol. 85, no. 8, pp. 46484650, 1999.
[26] J. F. Charpentier and G. Lemarquand, A comparative analysis of permanent magnet-type
bearingless synchronous motors for fully magnetically levitated rotors, Journal of Applied
Physics, vol. 83, no. 11, pp. 71217123, 1998.
[27] R. H. Frazier, P. J. Gilinson, and G. A. Oberbeck, Magnetic and Electric Suspensions, MIT
Press, 1974.
[28] J. Jin and T. Higuchi, Dynamics and stabilization of magnetic suspension using tuned lc
circuits, in Proceedings of the IV International Symposium on Magnetic Bearings, ETH,
Zurich, Switzerland, 1994.
[29] M. Tinkham, Introduction to Superconductivity, McGraw-Hill, Inc., New York, 2-nd edition,
1996.

BIBLIOGRAPHY

100

[30] D. Saint-James, G. Sarma, and E. J. Thomas, Type II Supeconductivity, Pergamon Press,


New York, 1969.
[31] A. Boerdijk, Levitation par des champs magnetostatiques, Rev. Tech. Philips, vol. 18, pp.
83, 1956.
[32] R. D. Waldron, Diamagnetic levitation using pyrolytic graphite, Rev. Sci. Instrum., vol. 37,
pp. 29, 1966.
[33] L. S. Wilk, Pseudodiamagnetic suspension, Rev. Sci. Instrum., vol. 43, no. 2, pp. 251258,
1972.
[34] V. M. Ponizovskii, Magnetic systems for suspending diamagnetic bodies freely, Pribory i
Tekhnika Eksperimenta, USSR, vol. 20, no. 5, pp. 178179, SeptemberOctober 1977.
[35] V. M. Ponizovskii, Diamagnetic suspension and its applications (survey),
Tekhnika Eksperimenta, USSR, vol. 24, no. 4, pp. 714, JulyAugust 1981.

Pribory i

[36] M. D. Simon and A. K. Geim, Diamagnetic levitation: Flying frogs and floating magnets
(invited), Journal of Applied Physics, vol. 87, no. 9, pp. 62006204, May 2000.
[37] J. R. Hull, Superconducting bearings, Superconductor Science & Technology, vol. 13, no.
2, pp. R1R15, 2000.
[38] V. Arkadiev, A floating magnet, Nature, vol. 160, pp. 330, 1947.
[39] I. Simon, Forces acting on superconductors in magnetic fields, Journal of Applied Physics,
vol. 24, pp. 19, 1953.
[40] P. K. Chapman and E. Ezekiel, Superconducting suspension for a sensitive accelerometer,
Rev. Sci. Instrum., vol. 36, pp. 96, 1965.
[41] F. Shiota, Hara K., and Hirata T., Improvement of the superconducting magnetic levitation
system for an absolute determination of the magnetic flux quantum, Japanese Journal of
Applied Physics, vol. 23, no. 4, pp. L227L229, 1984.
[42] A. Rivetti, G. Martini, R. Goria, and S. Lorefice, Turbine flowmeter for liquid helium with
the rotor magnetically levitated, Cryogenics, vol. 27, no. 1, pp. 811, 1987.
[43] R. Williams and J.R. Matey, Equilibrium of a magnet floating above a superconducting
disk, Applied Physics Letters, vol. 52, no. 9, pp. 751753, 1988.
[44] L.C. Davis, Lateral restoring force on a magnet levitated above a superconductor, Journal
of Applied Physics, vol. 67, no. 5, pp. 26312636, 1990.
[45] U. Schmitt, A. Tonoli, and H. J. Bornemann, Rotor dynamics of superconducting magnetic
bearings, Journal of Applied Physics, vol. 80, no. 2, pp. 12541256, 1996.
[46] J. Ohmori, H. Nakao, T. Yamashita, Y. Sanada, M. Shudou, M. Kawai, M. Fujita, M. Terai,
and A. Miura, Heat load tests of superconducting magnets vibrated electromagnetically for
the Maglev train, Cryogenics, vol. 37, no. 8, pp. 403407, 1997.

BIBLIOGRAPHY

101

[47] A. A. Kordyuk, Magnetic levitation for hard superconductors, Journal of Applied Physics,
vol. 83, no. 1, pp. 610612, 1998.
[48] C. K. McMichael, K. B. Ma, M. A. Lamb, M. W. Lin, R. L. Chow, L. Meng, P. H. Hor, and
W. K. Chu, Practical adaptation in bulk superconducting magnetic bearing applications,
Journal of Applied Physics, vol. 60, no. 15, pp. 18931895, 1992.
[49] J. R. Hull, J. L. Passmore, T. M. Mulcahy, and T. D. Rossing, Stable levitation of steel
rotors using permanent magnets and high temperature superconductors, Journal of Applied
Physics, vol. 76, no. 1, pp. 577580, 1994.
[50] J. R. Hull, E. F. Hilton, T. M. Mulcahy, Z. J. Yang, A. Lockwood, and M. Strasik, Low
friction in mixed-mu superconducting bearings, Journal of Applied Physics, vol. 78, no. 11,
pp. 68336838, 1995.
[51] T. A. Coombs and A. M. Campbell, Gap decay in superconducting magnetic bearings under
the influence of vibrations., Physica C, vol. 256, no. 34, pp. 298302, 1996.
[52] J. R. Hull and Mulcahy T. M., Mixed-mu superconducting bearings, US Patent 5 722 303,
March 3, 1998.
[53] C. K. McMichael and W. K. Chu, High temperature superconducting magnetic bearings,
US Patent 5 177 387, January 5, 1993.
[54] F. C. Moon and R. Raj, Superconducting rotating assembly, US Patent 4 886 778, December 12, 1989.
[55] A. V. Filatov and S. F. Konovalov, Low-hysteresis interaction of a hard type II superconductor with a permanent magnet, Physica C, vol. 271, no. 34, pp. 225229, 1996.
[56] A. V. Filatov, O. L. Poluschenko, and Sung-Chul Shin, A new approach to the design of
passive magnetic bearings using high-temperature superconductors, Cryogenics, vol. 38,
no. 6, pp. 595600, 1998.
[57] Sung-Chul Shin and A. V. Filatov, High-temperature superconducting magnetic bearing,
US Patent 5 789 837, August 4, 1998.
[58] Sung-Chul Shin and A. V. Filatov, High-temperature superconducting magnetic bearing,
Germany Patent 19641438, February 22, 2001.
[59] J. G. Bednorz and K. A. Muller, Possible high Tc superconductivity in the Ba-La-Cu-O
system, Zeitschrift fur Physik B-Condensed Matter, vol. 64, no. 2, pp. 189193, 1986.
[60] M. K. Wu, J. R. Ashburn, C. J. Torng, P. H. Hor, R. L. Meng, L. Gao, Z. J. Huang, Y. Q. Wang,
and C. W. Chu, Superconductivity at 93 K in a new mixed-phase Y-Ba-Cu-O compound
system at ambient pressure, Physical Review Letters, vol. 58, no. 9, pp. 908910, 1987.
[61] W. E. Smith, Electromagnetic levitation forces and effective inductance in axially symmetric
systems, Brit. J. Appl. Phys., vol. 16, pp. 377, 1965.
[62] E. R. Laithwaite, Electromagnetic levitation, Proceedings IEE, vol. 112, pp. 2361, 1965.

BIBLIOGRAPHY

102

[63] L. M. Holmes, Stability of magnetic levitation, Journal of Applied Physics, vol. 49, no. 6,
pp. 31023109, 1978.
[64] T. B. Jones, A necessary condition for magnetic levitation, Journal of Applied Physics,
vol. 50, no. 7, pp. 50575058, 1979.
[65] J. F. Bird, Theory of magnetic levitation for biaxial systems, Journal of Applied Physics,
vol. 52, no. 2, pp. 578588, 1981.
[66] M. Bonvalot, P. Courtois, P. Gillon, and R. Tournier, Magnetic levitation stabilized by eddy
currents, Journal of Magnetism & Magnetic Materials, vol. 151, no. 12, pp. 283289,
1995.
[67] R. A. Clemente and M. Tessarotto, Static plasma confinement by harmonic magnetic fields,
Journal of the Physical Society of Japan, vol. 67, no. 1, pp. 163165, 1998.
[68] R. A. Clemente, Minimum temperature in electromagnetic levitation melting under terrestrial gravity, Journal of Applied Physics, vol. 84, no. 5, pp. 29682970, 1998.
[69] M. Enokizono and H. Tanabe, Numerical analysis of high-frequency induction heating
including temperature dependence of material characteristics, IEEE Transactions on Magnetics, vol. 31, no. 4, pp. 24382444, 1995.
[70] H. Tadano, M. Fujita, T. Take, K. Nagamatsu, and A. Fukuzawa, Levitational melting of
several kilograms of metal with a cold crucible, IEEE Transactions on Magnetics, vol. 30,
no. 6, pp. 47404742, 1994.
[71] K. A. Connor and J. A. Tichy, Analysis of an eddy current journal bearing, Transactions
of the ASME, vol. 110, pp. 321326, 1988.
[72] Leehua Ting and John Tichy, Stiffness and damping of an eddy current magnetic bearing,
Journal of Tribology, vol. 114, pp. 600605, 1992.
[73] F. R. Post,
Passive magnetic bearing element with minimal power losses, US
Patent 5 847 480, December 8, 1998.
[74] James R. Powell Jr. and Gordon T. Danby, Electromagnetic inductive suspension and stabilization system for a ground vehicle, US Patent 3 470 828, October 7, 1969.
[75] K. R. Davey, Designing with Null Flux Coils, IEEE Trans. Magn., vol. 33, no. 5, pp.
43274334, 1997.
[76] K. Sawada, Development of magnetically levitated high speed transport system in Japan,
IEEE Trans. Magn., vol. 32, no. 4, pp. 22302235, 1996.
[77] M. Ono, S. Koga, and Ohtsuki H., Japans superconducting Maglev train, IEEE Inst. and
Meas. Mag., vol. 5, no. 1, pp. 915, 2002.
[78] J. L. He, D. M. Rote, and H. T. Coffey, Applications of the dynamic circuit theory to Maglev
suspension systems, IEEE Trans. Magn., vol. 29, no. 6, pp. 41534164, 1993.

BIBLIOGRAPHY

103

[79] J. L. He and H. T. Coffey, Magnetic damping forces in figure-eight-shaped Null-Flux Coil


suspension systems, IEEE Trans. Magn., vol. 33, no. 5, pp. 42304232, 1997.
[80] K. Davey, T. Morris, J. Schaff, and Rote D., Calculation of motion induced eddy current
forces in Null Flux Coils, IEEE Trans. Magn., vol. 31, no. 6, pp. 42144216, 1995.
[81] K. Davey, Electrodynamic Maglev coil design and analysis, IEEE Trans. Magn., vol. 33,
no. 5, pp. 42274229, 1997.
[82] K. Davey, Analysis of an electrodynamic Maglev system, IEEE Trans. Magn., vol. 35, no.
5, pp. 42594267, 1999.
[83] Final Report on the National Maglev Initiative, US Department of Transportaton, Federal
Railroad Administration, 1993.
[84] W. A. Jacobs, Magnetic launch assist NASAs vision for the future, IEEE Trans. Magn.,
vol. 37, no. 1, pp. 5557, 2001.
[85] J. F. Pinkerton, Magnetic bearing and method utilizing movable closed conductive loops,
US Patent 5 302 874, April 12, 1994.
[86] J. F. Pinkerton, D. B. Clifton, and R. L. Scott, Magnetic bearing with closed loops which
move in AC electromagnet field and/or which include capacitors, US Patent 5 345 128,
September 6, 1994.
[87] D. B. Clifton, R. L. Scott, and J. F. Pinkerton, Null flux magnetic bearing with crossconnected loop portions, US Patent 5 471 105, November 28, 1995.
[88] J. A. Andrews and J. F. Pinkerton,
Patent 5 508 573, April 16, 1996.

Magnetic bearing with phase-shifted loops, US

[89] F. R. Post, Dynamically stable magnetic suspension/bearing system, US Patent 5 495 221,
February 27, 1996.
[90] R. M. Harrigan, Levitation device, US Patent 4 382 245, May 3, 1983.
[91] E. W. Hones and W. G. Hones,
Patent 5 404 062, April 4, 1995.

Magnetic levitation device and method, US

[92] M. D. Simon, L. O. Heflinger, and S. L. Ridgway, Spin stabilized magnetic levitation,


American Journal of Physics, vol. 65, no. 4, pp. 286292, 1997.
[93] M. V. Berry, The Levitron: an adiabatic trap for spins, Proc. Roy. Soc., Ser. A, vol. 452, no.
1948, pp. 12071220, May 1996.
[94] R. F. Gans, T. B. Jones, and M. Washizu, Dynamics of the Levitron, Journal of Physics,
D-Applied Physics, vol. 31, no. 6, pp. 671679, 1998.
[95] S. Gov, S. Shtrikman, and H. Thomas, On the dynamical stability of the hovering magnetic
top, Physica D, vol. 126, no. 34, pp. 214224, 1999.
[96] P. A. Simpson, Damper system for suspension systems, US Patent 3 929 390, December 30,
1975.

BIBLIOGRAPHY

104

[97] E. J. Gunter, R. R. Humphris, and S. J. Severson, Design study of magnetic eddycurrent vibration dampers for application to cryogenic turbomashinery, Tech. Rep.
UVA/528210/MAE84/101, NASA Grant NAG-3-263, University of Virginia.
[98] K. Nagaya, H. Kojima, Y. Karube, and H. Kibayashi, Braking forces and damping coefficients of eddy-current brakes consisting of cylindrical magnets and plate conductors of
arbitrary shape, IEEE Trans. Magn., vol. 20, no. 6, pp. 21362145, 1984.
[99] J. R. Frederick and M. S. Darlow, Operation of an electromagnetic eddy-current damper
with a supercritical shaft, Trans. ASME, Journal of Vibration and Acoustics, vol. 116, no. 6,
pp. 578580, 1994.
[100] G. Genta, C. Delprete, A. Tonoli, E. Rava, and L. Mazzocchetti, Analytical and experimental
investigation of a magnetic radial passive damper, in Proceedings of the Third International
Symposium on Magnetic Bearings, 1992, vol. 116, pp. 255264.
[101] Y. Kligerman, A. Grushkevich, and M. S. Darlow, Analytical and experimental evaluation
of instability in rotordynamic system with electromagnetic eddy-current damper, Trans.
ASME, Journal of Vibration and Acoustics, vol. 120, pp. 272278, 1998.
[102] J. P. Den Hartog, Mechanical Vibrations, 3rd edition.
[103] A. V. Filatov and E. H. Maslen, A passive magnetic bearing, in Proceedings of the VII
International Symposium on Magnetic Bearings, ETH, Zurich, Switzerland, 2000.
[104] A. V. Filatov and E. H. Maslen, High speed, high efficiency magnetic bearing, in Proceedings of the VI International Symposium on Magnetic Suspension Technology, Turin, Italy,
2001, pp. 5863.
[105] A. V. Filatov and E. H. Maslen, Passive magnetic bearing for flywheel energy storage
systems, IEEE Trans. Magn., vol. 37, no. 6, pp. 39133924, 2001.
[106] A. V. Filatov, E. H. Maslen, and G. T Gillies, A method of noncontact suspension of rotating
bodies using electromagnetic forces, J. Appl. Phys., vol. 91, no. 4, pp. 23552371, 2002.
[107] Filatov A. V. and Salter A. K., Magneto-dynamic bearing, US Patent 6 304 015, October 16,
2001.
[108] Heinz Knoepfel, Magnetic Fields, John Wiley & Sons, Inc., 2000.
[109] John Jackson, Classical Electrodynamics, John Wiley & Sons, Inc., 3-rd edition, 1999.
[110] Richard Dorf and Robert Bishop, Modern Control Systems, Addison-Wesley, 7-th edition,
1995.
[111] D. C. Meeker and E. H. Maslen, Prediction of rotating losses in heteropolar radial magnetic
bearings, Journal of Tribology, vol. 120, no. 3, pp. 629635, 1998.
[112] R. L. Stoll, The analysis of eddy currents, Oxford University Press, London, 1974.
[113] Hassan K. Khalil, Nonlinear Systems, Prentice Hall, Upper Saddle River, 2-nd edition, 1996.

Appendix A

MathCAD script for calculating the


exact stability bounds.

105

APPENDIX A. EXACT STABILITY BOUNDS

106

Appendix B

Example of the force calculation:


conductor G is shaped as a set of
separate loops.
In this illustrative example we calculate forces acting on the rotor including n conducting loops
such as shown in Figure 2.3, which is repeated here for convenience as Figure B.1. This example
will clarify the physical meanings of the parameters K and and give practically useful equations
to calculate these parameters for this particular type of the rotor.
For simplicity we assume that there is no inductive and/or resistive coupling between the loops.
Because of this we can consider each loop separately, find forces acting on it and then sum forces
acting on all the loops. We assume that the rotor is displaced by a fixed distance r0 .

B.1

Direct force calculation

First, we calculate forces acting on the loops directly. Figure B.2 shows one loop rotating with an
angular speed about the axis Z displaced from the field symmetry axis Z0 by a distance r0 . We
start timing when the middle line of this loop coincides with the displacement direction.
Let us find the change of the magnetic flux through the loop during the rotation. From the

Figure B.1: The operational principle of an embodiment of the proposed suspension.

107

APPENDIX B. FORCE CALCULATIONS

108

Figure B.2: Direct calculation of the forces acting on a conducting loop.


triangle Z0 ZA:
R22 = (ZA)2 + 2r0 + 2(ZA)r0 cos 1
Solving this equation about (ZA) we get

(ZA) =

2r0 cos 1 2


R22 r2 (1 cos2 1 )
R2 r cos 1
2

The elementary area shown dashed on Figure B.1 is


dS2 =


1 2
R2 (R2 r0 cos 1 )2 d1 R2 r0 cos 1 d1 .
2

Therefore, the change of the area under the outer arc of the conducting loop is

t+0
R2 r0 cos 1 d1 = R2 r0 (sin(t + 0 ) sin(t 0 )) =
S2 =
t0

= 2R2 r0 sin 0 cos t.


By analogy, the change of the area under the inner arc is
S1 = 2R1 r0 sin 0 cos t.
The change of the external magnetic flux through the loop is
ext = B(S2 S1 ) = 2B(R1 + R2 )r0 sin 0 cos t = 4BRav r0 sin 0 cos t,

APPENDIX B. FORCE CALCULATIONS

109

where
Rav =

R1 + R2
.
2

Differentiating yields
dext
= 4BRav r0 sin 0 sin t.
dt
The current in the loop is given by the following equation:
L

dext
dI
+ RI =
= 0 sin t,
dt
dt

where
0 = 4BRav r0 sin 0 .
In the steady mode, the current is
I = I0 cos(t + );

I0 =

4BRav sin 0

r0 ;
L 1 + tan2

= arctan

R
,
L

(B.1)

where R and L are the resistance and the inductance of the loop respectively.
The force acting on the loop in the displacement direction (r-direction) is

t+0
IB(R2 + R1 ) cos 1 d1 = 2IBRav (sin(t + 0 ) sin(t 0 )) =
Fr1 =
t0

= 4IBRav sin 0 cos t.


Substituting equation (B.1) for current yields
Fr1 = F0 cos t cos(t + )
or
Fr1 =

F0
(cos + cos(2t + )) ;
2

F0 =

(4BRav sin 0 )2

r0 .
L 1 + tan2

(B.2)

The time invariant component of Fr1 ( the same as time averaged component) is
Favr1 =

(4BRav sin 0 )2 cos

r0
2L 1 + tan2

(B.3)

The force acting on the loop in the direction perpendicular to the displacement direction (tangential
direction) is

t+0
F 1 =
IB(R2 + R1 ) sin 1 d1 = 2IBRav ((cos(t + 0 ) cos(t 0 )) =
t0

= 4IBRav sin 0 sin t.


F 1 = F0 sin t cos(t + )

APPENDIX B. FORCE CALCULATIONS

110

or
F 1 =

F0
(sin sin(2t + )) ;
2

F0 =

(4BRav sin 0 )2

r0 .
L 1 + tan2

(B.4)

The time invariant (time averaged) component of F 1 is


Fav 1 =

(4BRav sin 0 )2 sin

r0 .
2L 1 + tan2

(B.5)

Assume that we have n loops located uniformly around the rotor axis Z. In terms of Section 3.1,
the rotor in this case has a rotational periodicity of the order n. As was proven in the Section 3.1,
whenever the rotating conductor has rotational periodicity of order higher than 3, the time variant
force components cancel each other. (In this case the rotational periodicity is represented by the
number of loops). This can be verified by direct calculation for a particular n. For this purpose, we
number all the loops in the direction of rotation starting with the loop considered before. Then we
to the
can rewrite force equations (B.2) and (B.4) for the k-th loop by adding a phase shifts 2(k1)
n
angle t. It is easy to verify for a particular n by direct calculation that
n

k=1

(k 1)
sin 2t + +
n


=

n

k=1



(k 1)
=0
cos 2t + +
n

Then the components of the total force produced by all n loops are
Fr =

n (4BRav sin 0 )2 cos

r0
2
L 1 + tan2

(B.6)

F =

n (4BRav sin 0 )2 sin

r0 .
2
L 1 + tan2

(B.7)

The magnitude of the total force is

F =


n (4BRav sin 0 )2

r0
Fr2 + F2 =
2 L 1 + tan2

(B.8)

From the equations (B.6) and (B.7), it is easy to see that this force is deflected from the direction
opposing the displacement by the angle . Since this is the definition of the angle given in the
Section 3.1, we conclude that = . Thus, when the conductor G is formed by n independent
conducting loops with inductance L and resistance R, the angle is


R
.
(B.9)
= arctan
L
From the equation (B.8), it is easy to see that the in-plane stiffness K also introduced in Section 3.1 in this case is
K=

n (4BRav sin 0 )2

.
2 L 1 + tan2

(B.10)

APPENDIX B. FORCE CALCULATIONS

111

Finally, we find the drag torque acting on the rotor (neglecting the eddy currents induced in the
walls of the conducting loops). As can be seen from Figure B.2, forces acting on the arcs of the
conducting loops are directed radially and cannot cause any drag torque. The forces exerted on the
radial parts of a loop are
Ft1 = 2Br0 I cos(t + 0 );
Ft2 = 2Br0 I cos(t 0 ).
Therefore, the drag torque due to one loop is
T1 = (Ft1 + Ft2 )Rav = 4Br0 I sin 0 sin t,
1
T1 = T0 cos(t ) sin t = T0 (sin + sin(2t )),
2
where
T0 =

(4BRav sin 0 )2 2

r0 .
L 1 + tan2

Summing the torque values acting on all the loops, we find again that the oscillating torque components cancel each other and the total value of the drag torque acting on the rotor is
T =

n (4BRav sin 0 )2 R 2
n (4BRav sin 0 )2 sin 2

r0 =

2
2 L2 (1 + tan2 ) r0
L 1 + tan2

(B.11)

Note that
T = F r0 sin
and this expression could be obtained much more easily using Lemma 1 from Section 3.1.
The power loss during the rotation is
Prot = T =

n (4BRav sin 0 )2 R 2
r0 .
2 L2 (1 + tan2 )

If we calculate the resistive power loss due to the currents flowing in the loop, we obtain exactly
the same value:
Pres =

B.2

n 2
n (4BRav sin 0 )2 R 2
I0 R =
r0 .
2
2 L2 (1 + tan2 )

Force calculation using energy methods (Section 3.1)

Time averaged radial and tangential force components acting on an arbitrary rotating conductor
G can be obtained from the general equations (3.30) and (3.35) derived using energy methods.
Considering that we neglected inductive and resistive couplings between the coils, all the matrices
are diagonal and the total force produced by n coils is simply n times the time-averaged force due
to one coil. Therefore, it is sufficient to carry out the calculation only for one coil, in which case all
the matrices become scalars:
LG = L;

RG = R;

= I0 cos(t + ) (see Eq. (B.1)).

APPENDIX B. FORCE CALCULATIONS

112

It is easy to see that


< 2 >=

I02
(4BRav sin 0 )2
=
.
2
2L2 (1 + tan2 )

Then for one coil the equations (3.30) and (3.35) yield
Fav r1 = L < 2 > r0 =

(4BRav sin 0 )2
r0
2L (1 + tan2 )

R
(4BRav sin 0 )2 R
< 2 > r0 =
r0 .

2L2 (1 + tan2 )
0.5

= cos , one can see easily that this equations are identical to
Noting that 1 + tan2
(B.3) and (B.5).
The equation (3.36) readily gives us


R
= .
= arctan
L
Fav 1 =

(The same as equation (B.9)).

Appendix C

Calculation of the magnetic field in the


radial suspension system.
To generate the magnetic field in the radial suspension system, we consider the structure shown in
Figure C.1. We analyze two cases: when the structure shown in Figure C.1 includes pole shoes
covering the magnet surfaces facing the air gap, and when it does not.
The variant with pole shoes potentially allows generating a field, which is more uniform in the
circumferential direction. The field uniformity is in large degree insensitive to the uniformity of
the magnet properties, in particular circular magnets can be assembled from separate blocks as it is
done in the prototype described in this dissertation. Furthermore, as will be shown soon, this variant
features wider field plateaux (areas of low field gradients) in the radial directions, thus allowing for
wider conducting arcs. A possible disadvantage is an increase of the rotating coil inductance L ,
which causes reduction of the radial suspension stiffness K (see equation (B.10) at high rotational
speeds, when the angle is close to zero and does not depend much on L. At the same time,
an increase of L reduces at each particular speed, resulting in a lower minimal leviation speed
achievable with a given damping coefficient.
While it is possible to model magnetic fields more accurately, the goal of the following analysis
is obtaining simple analytical equations, which will give insights into how different system parameters influence the field distribution. These equations are also useful in estimating bounds on the
load characteristics achievable in this system.

Figure C.1: A variant of the magnetic system used to generate a magnetic field as shown in Figure 4.4.

113

APPENDIX C. MAGNETIC FIELD CALCULATIONS

114

Figure C.2: A model for the field calculation in the system without pole shoes.

C.1 Magnetic system without pole shoes.


First, we calculate the magnetic field distribution (more accurately the distribution of the axial
field component, which is of primary interest since it is the only component which produces the
radial Lorenz force) for the case not including pole shoes. Modern magnetic materials, including
NdFeB and ferrites, have residual magnetization M0 essentially independent of the demagnetizing
magnetic field [2]. The magnetic field produced by such magnets can be calculated by introducing
a fictitious surface density of the magnetic charge m defined as
m = (0 M0 )n ,
where (0 Mn )0 is the component of the magnetic polarization normal to the magnet surface. Then,
assuming that the magnets in Figure C.1 are thick enough so that we can neglect influence of the
back iron on the field in the air gap, the latter can be calculated as if generated by the fictitious
magnetic charges on the magnet surfaces facing the gap using the inverse square law. (If necessary,
the back iron can be taken into account by specifying zero potential on its surface.)
The method of the magnetic field calculation, which we will use here, is based on the similarities
between the following equations for the electrical and magnetic fields:
In the vacuum:
E = 0;

D =0

for the electrical field

H = 0;

B =0

for the magnetic field.

In a medium:
D = 0 E + Pe

for the electrical field

B = 0 H + 0 M

for the magnetic field.

The magnetic field in the system shown in Figure C.1 can, therefore, be calculated as a field
produced by four layers with surface charge densities +0 M and 0 M as shown in Figure C.2.
It is easy to show that the z-component of the magnetic field produced by a charged rectangle
x [b; b]; y [l; l] is
1
m
Bz =
4

dy
l

b (x2

z
+

y2

+ z 2 )3/2

dx

APPENDIX C. MAGNETIC FIELD CALCULATIONS

115

This integral can be evaluated precisely. The final expression is




(b x)(l y)
0 M

Bz =
arctan
+
4
z (b x)2 + (l y)2 + z 2


(b + x)(l y)
0 M

arctan
+
+
4
z (b + x)2 + (l y)2 + z 2


(b x)(l + y)
0 M

arctan
+
+
4
z (b x)2 + (l + y)2 + z 2


(b + x)(l + y)
0 M

arctan
.
+
4
z (b + x)2 + (l + y)2 + z 2
If l , then
0 M
Bz =
2






bx
b+x
arctan
+ arctan
z
z

(C.1)

We will use this approximation to calculate magnetic fields in our rotational system, assuming that
it can be unrolled.
The axial magnetic field produced by the system shown in Figure C.1 can be then calculated as
a superposition of the fields produced by the four layers:





x /2
x + 2b + /2
0 M
arctan
+ arctan

Bz =
4
z g/2
z g/2





x /2
x + 2b + /2
0 M
arctan
+ arctan

4
z + g/2
z + g/2





0 M
x /2
x + 2b + /2

arctan
+ arctan
+
4
z g/2
z g/2





x /2
x + 2b + /2
0 M
arctan
+ arctan
.
+
4
z + g/2
z + g/2
In the middle of the air gap (z = 0), the magnetic field is





x + /2
x /2
0 M
arctan
+ arctan

Bz0 =
2
g/2
g/2





x + 2b + /2
x 2b /2
0 M
arctan
+ arctan
.

4
g/2
g/2
It is convenient to describe the field in dimensionless coordinates
x
=

x
.
b

We also dimensionalize the geometric parameters of the system,g and :


g =

g
2b

= .
and
2b

APPENDIX C. MAGNETIC FIELD CALCULATIONS

116

Bz / Polarization

0.5

-0.5

-1

-2

-1

x/b

Figure C.3: Field distribution in the middle of the air gap calculated using constant charge density
= 1/3 and g = 0.4.
model with

Figure C.4: A model for the field calculation in the system with pole shoes.
With these designations, the magnetic field in the middle of the air gap is
/
1
2
1
20

x
+
x

0 M
arctan
+ arctan

Bz0 =
2
g
g
/
1
2
1
20

x
+2+
x
2
0 M
arctan
+ arctan
.

2
g
g

(C.2)

which results in a smooth transition (a nearly uniform field


By inspection, the highest value of
resulted in smaller
gradient) between two field plateaux was found to be 1/3. The lower values of
plateaux, higher - in unnecessarily large distances between the plateaux. The field distribution
= 1/3 and g = 0.4 is shown in Figure C.3.
calculated for

C.2 Magnetic system with pole shoes.


Next, we consider the system with pole shoes. In this case, the field can be calculated as if it is
produced by four conducting surfaces with potentials +0 and 0 (Figure C.4).
To simplify calculation, we assume that the equipotential surfaces also cover the gap and
we make a periodic expansion of the boundary potential as shown in Figure C.5. This results in a
boundary condition of the first kind (x, g/2) = (x, g/2) = b , which is easy to deal with.

APPENDIX C. MAGNETIC FIELD CALCULATIONS

117

Figure C.5: Periodic expansion of the system with pole shoes.


Here

if W (2n 1) < x < 2nW


0
0
if x = nW
b =

0 if 2nW < x < (2n + 1)W

(C.3)

and
W = 2b + .
Apparently, the solution in the case of the expanded system shown in Figure C.5 will differ from
the one in the case of the original system (Figure C.4). However, for practical values of g (0.5 and
lower), the major difference occurs outside of the magnet areas, where the field distribution is of no
interest.
The magnetic potential in the air gap is a solution of Laplaces equation:
2 2
+ 2 = 0.
x2
z
Because the system has odd periodicity in the x-direction, we look for a solution in the form of
a Fourier sine expansion over x:
(x, z) =

Gn (z) sin

n=1

n
x
W

(C.4)

Writing Laplaces equation for the n-th component, we get


d2 Gn ) n *2

Gn = 0
dz 2
W

(C.5)

To satisfy Gn (0) = 0 x, we need Gn in form


) n *
z .
Gn = An sinh
W
Thus
(x, z) =

An sinh

n=1

) n *
) n *
z sin
x
W
W

At z = g/2, we have
b (x) =


n=1

An sinh

) n g *
W 2

sin

) n *
x
W

(C.6)

APPENDIX C. MAGNETIC FIELD CALCULATIONS


Multiplying both the left and the right parts of the (C.6) by sin
W and using the orthogonality of sines we get
/
0
4
m sinh( m g ) if m = 2i + 1
W 2
Am =
0
if m = 2i

118
 m 
W x , integrating from W to

where i = 0, 1, 2, 3 . . ..
Summing up,
(x, z) =


i=0

sinh

(2i+1)
z
W

40
)
* sin
(2i + 1) sinh (2i+1) g
W


(2i + 1)
x
W

The z-component of the magnetic flux density then


)
*


cosh (2i+1) z

W
40
(2i + 1)
40
* sin
)
=
x
Bz (x, z) =
z
W
W
sinh (2i+1) g
i=0

(C.7)

At z = 0 we have

)
*
sin (2i+1) x

W
40
)
*.
Bz0 (x) = Bz (x, 0) =
(2i+1)
g
W
sinh
i=0

(C.8)

we rewrite (C.8) as
Noting that W = 2b + /2 = (2 + )b,
)
*
sin (2i+1) x

) 2+ * ,
Bz0 (x) = B0
(2i+1)

i=0 sinh
g

(C.9)

2+

where
B0 =

40
.
W

To express B0 in terms of the magnet thickness lpm and polarization 0 M, we find the magnetic
field at one point (x = W/2) and assume that the gap is narrow enough (
g is small) so that the
magnetic field is essentially uniform in the gap, directed along the z-axis and equal to some value
Bg . For x = W/2 we have
*
)

sin (2i+1)


2
(1)i
)
*=
)
* 1.4.
(2i+1)
(2i+1)
g

sinh
g

i=0 sinh
i=0
(2+e)
(2+e)
Therefore,
Bg = 1.4B0 .

(C.10)

Applying H = 0 to our magnetic system (Figure C.1), we obtain


Hg g = Hpm lpm ,

(C.11)

APPENDIX C. MAGNETIC FIELD CALCULATIONS

119

where Hg and Hpm are the magnetic field strengths in the gap and in the magnet. Inside the magnet,
we also have
Bpm = 0 M 0 Hpm ,

(C.12)

where Bpm is the flux density inside the magnet. All the field are assumed to be collinear. Strictly
speaking, 0 M is a function of Hpm , however, for modern magnetic materials it is essentially
constant. In addition, the flux continuity requires
Bpm = Bg .

(C.13)

Solving (C.11), (C.12) and (C.13) simultaneously, we find


Bg = (0 M)

lpm
.
g + lpm

(C.14)

Using (C.10), we find


B0 =

0.714 lpm
0.714 lpm
(0 M) =
(0 M).
g + lpm
g + lpm

(C.15)

where
lpm = lpm .
2b
Thus, finally we have
)
*
sin (2i+1) x

2+e
0.714 lpm
)
*,
(0 M)
Bz0 (x) =
(2i+1)

g + lpm
sinh
g
i=0

(C.16)

(2+e)

The first six terms of this series expansion already give a good approximation of the field distribution, which is shown in Figure C.6. Note that the field distribution shown there exhibits much
flatter and wider plateaux than in Figure C.3. The width W5% , where the field deviations from the
maximal value do not exceed 5%, were equal to 0.85 b and 1.43 b in the case of Figures C.3 and C.6
respectively.
Figure C.7 shows how constant-potential and constant-charge models approximate the field distribution in the gap of the bearing prototype. The areas of negative and the positive fields were not
exactly mirror images of each other, because of the different amounts of magnetic material used. To
make the field distribution symmetric about the X axis, the field pattern was shifted up by 0.012 T .
The magnetic field distribution given by equation (C.16) gives a good fit in the area of the air gap.
Outside of the gap, the model fails because of the periodic extension assumption.

C.3 Choosing parameters of the magnetic system.


One of the important parameters of the magnetic system is the thickness of the permanent magnets
lpm (Figure C.1). We choose this parameter in order to maximize the ratio of the force acting on
a conducting loop displaced from equilibrium by a certain amount (or the suspension stiffness) vs.

APPENDIX C. MAGNETIC FIELD CALCULATIONS

120

Bz / Polarization

0.5

-0.5

-1

-2

-1

x/b

Figure C.6: Field distribution in the middle of the air gap calculated using constant potential model
= 1/3 and g = 0.4.
with

Bz / Polarization

0.5

-0.5

-1

-2

-1

x/b

Figure C.7: Comparison of the field distributions obtained with constant charge and constant potential models with the one measured in the prototype magnetic system employing pole shoes.

APPENDIX C. MAGNETIC FIELD CALCULATIONS


Parameter
Pole width, 2b
Axial air gap, g
Radial air gap,
Magnet thickness, lpm
Back iron thickness, lF e
Field uniformity area, W5%
Total width, W = 2(2b) +
Total height, H = lpm + lF e + g

Dimensionless representation
2b/(2b)
g = g/(2b)
= /(2b)

lpm = lpm /(2b)


lF e = lF e /(2b)
5% = W5% /(2b)
W
= W/(2b)
W
= H/(2b)
H

121
Dimensionless value
1
0.4
1/3
0.4
0.3
0.72
7/3
1.1

Table C.1: Suggested geometry of the radial suspension magnetic system with pole shoes. Parameters are calculated assuming 0 M = 1.2 T (NdFeB magnets) and Bs = 2 T (iron).

the thickness of the permanent magnets. It is easy to see that this force (stiffness) is proportional
to the second power of the magnetic flux density in the air gap (see, for example, equation (B.10)).
Thus we need to maximize the ratio of the second power of the magnetic field in the gap vs. the
magnet thickness. Noting that the field energy density is also proportional to the second power of
the flux density, we can say that we maximize the energy of the field generated by the magnets vs.
the magnet volume.
From the equation (C.14), we find
Bg2
lpm
= (0 M)2
.
lpm
(g + lpm )2

(C.17)

Taking the derivatives of the right-hand part with respect to lpm , we find that the ratio on the lefthand sides reaches a maximum when lpm = g. In this case Bg = 0 M/2.
The next parameter we need to find is the combined thickness of the back iron pieces, lF e
(Figure C.1). The minimal value of this parameter is dictated by the ability of these iron pieces to
carry the flux without saturation. We assume that the saturation field is Bs and that the field in the
gap is uniformly distributed over the magnet poles with the field density being Bg . Then, the flux
continuity gives us
2bBg = Bs lF e ,
from which we find
lF e =

Bg
2b.
Bs

(C.18)

In the Table C.1 , we summarize all the geometrical parameters of the radial suspension magnetic
system, which are calculated for g = 0.4 (this is the characteristic number of the magnetic system
used in the presented bearing prototype).

Appendix D

Estimations of the potential loading


characteristics of the radial suspension.
The most critical part of the proposed bearing is the electrodynamic radial suspension. In particular,
the stiffness produced by this suspension defines the sum of the stiffnesses achievable in all three
lateral directions in accordance with the equation (1.3), which is repeated here for convenience:
Kax + 2Krad 0.
In this appendix we estimate maximal stiffness and other loading characteristics of the radial
suspension. Note that we consider only static loading characteristics resulting from the interaction
of a rotating conductor with a stationary magnetic field. We do not consider the damping system
it is assumed that we can always achieve necessary damping (if necessary, an active damper
can be used). Consequently, when we calculate, for example, the rotor volume or weight, the
volume (weight) of the damping system is neglected, and such bearing parameters as load capacity
or stiffness per unit volume or weight are overestimated. It can be expected, however, that the
weight of the damping system would not exceed the weight of the positioning system, and the
numbers can be corrected accordingly (e.g., the ratio of the load capacity vs. bearing volume with
a damper would be approximately half of that calculated without the damper). It is also believed
that if an active damper is used, its volume and weight would be negligible compared to those of
the positioning subsystem.
It is further to be noted that we do not perform a careful parameter optimization our estimates
of the achievable performance are based on the design parameters which we selected for the bearing
prototype. A better performance can be obtained through an optimization procedure.
The major parameters of the prototype needed for the loading characteristic calculation are
summarized in Table D.1.
Figure D.1 indicates how much stiffness per radial footprint area of the bearing could be obtained in the current prototype if NdFeB magnets were used instead of the ceramic ones and the
additional inductors were eliminated. The stiffness is calculated using equation (B.10) and the radial footprint area is defined as Sf p = 2Rout H, where Rout = Rav + W/2 is the outer radius of
the bearing; Rav is the average radius of the coil (see Table D.1), or the magnetic system; W and
H are the total width and height of the magnetic system respectively (see Table C.1). Thus, in the
prototype we have Rout =167 mm and, if we replace currently used ceramic magnets with NdFeB,
the optimal height of the magnetic system will be H=2.5 cm. (Currently used H=3.94 cm is not
122

APPENDIX D. LOADING CHARACTERISTICS

123

Parameter
Number of coils, n
Average radius, Rav
Coil angle, 0
Width of the coil walls, Wc
Height of the coil walls, Hc
Distance between the middle lines of the coil walls, a
Equivalent inductance of a coil*, Le
Equivalent resistance of a coil, Re

Value
4
140 mm
34.50 0.6 rad
1.15 cm
0.5 cm
2.8 cm
1.5 107 H
1.6 104

Table D.1: Parameters of the rotating coils in the bearing prototype.


* - The inductance of the coil was calculated assuming a uniform current distribution through the
coil cross-section. It does not include the effect of the pole shoes.

Parameter
Pole width*, 2b
Average magnetic flux density, Bav

Value
2.3 cm
0.13 T

Table D.2: Parameters of the stationary magnetic system used in the radial suspension.
* - The other sizes can be calculated using Table D.1

Stiffness/Footprint, kPa/mm

60
50
40
30
20

Current sizes
Current sizes x2

10

100

200

300

400

Rotational frequency, Hz

Figure D.1: Predictions of the stiffness per radial footprint area achievable with the sizes of the
current prototype and when the sizes are scaled up by a factor of two.

APPENDIX D. LOADING CHARACTERISTICS

124

Load Capacity/Footprint, kPa

280
240

Current sizes
Current sizes x2

200
160
120
80
40
100

200

300

400

Rotational frequency, Hz

Figure D.2: Predictions of the load capacity per radial footprint area achievable with the sizes of the
current prototype and when the sizes are scaled up by a factor of two.
Material
NdFeB
Cu
Fe

Density, g/cm3
8.1
8.9
7.8

Table D.3: Densities of the major materials used in the radial bearing.

optimal; its choice is dictated by the height of commercially available ceramic magnet blocks). Note
that the stiffness-per-surface-area saturation levels do not depend on the scaling factor.
Given the width of the conducting loop wall, Wc , (see Table D.1) we can find the load capacity
of the bearing, assuming that the wall does not leave the area of the field plateau. We will use
W5% (see Table C.1) as a measure of the plateau width. This implies that the field variation in the
conducting wall will not exceed 5% of the field amplitude. Then the maximal radial displacement
of the rotor is max = (W5% Wc )/2. In the current prototype, max =2.5 mm, but it is further
limited to 2 mm by the start-up bearings.
Figure D.2 shows load capacity per radial footprint area achievable with the sizes of the current
prototype and when the sizes are scaled up by a factor of two. We assumed that max =2.5 mm.
Note that the saturation level of the load capacity per surface area goes up proportionally to the
scaling factor (stiffness does not change and the maximal displacement increases).
We can also estimate achievable
 2 stiffness
 and load capacity per unit volume of the bearing. The
2
latter is calculated as V = Rout Rin H, where Rin = Rav W/2 is the inner radius of the
bearing. Such estimates are shown in Figures D.3 and D.4.
Finally, such parameters as stiffness and load capacity per weight of the bearing or per weight of
the rotor may be important, especially for space applications. To estimate the weights of the bearing
components, we need the densities of NdFeB, copper and iron. These are provided in Table D.3.
Note that the weights of the bearing and the rotor used in these calculation are, in fact, the weights
without any supplementary components. Thus, we use the weight of the four copper coils as the
weight of the rotor. The coil volume in the prototype is 21.4 cm3 , as calculated by AutoCAD using
the existing drawing. The mass of one coil is 190 g; the mass of four coils is consequently 760 g,

APPENDIX D. LOADING CHARACTERISTICS

125

Stiffness/Volume, N/cm

3.5
3
2.5
2
1.5

Current sizes
Current sizes x2

1
0.5
100

200

300

400

Rotational frequency, Hz

Load Capacity/Volume, N/cm

Figure D.3: Predictions of the stiffness vs. bearing volume ratios achievable with the sizes of the
current prototype and when the sizes are scaled up by a factor of two.

0.5

Current sizes
Current sizes x2

100

200

300

400

Rotational frequency, Hz

Figure D.4: Predictions of the load capacity vs. bearing volume ratios achievable with the sizes of
the current prototype and when the sizes are scaled up by a factor of two.

Stiffness/Bearing weight, 1/mm

APPENDIX D. LOADING CHARACTERISTICS

126

Current sizes
Current sizes x2

100

200

300

400

Rotational frequency, Hz

Load Capacity/Bearing weight

Figure D.5: Predictions of the stiffness vs. bearing weight ratios achievable with the sizes of the
current prototype and when the sizes are scaled up by a factor of two.

20

15

10

Current sizes
Current sizes x2

100

200

300

400

Rotational frequency, Hz

Figure D.6: Predictions of the load capacity vs. bearing weight ratios achievable with the sizes of
the current prototype and when the sizes are scaled up by a factor of two.
and the weight of the four coils is 7.5 N. Calculations of the volumes, masses and weights of the
stationary components, all of which have cylindrical shapes, are elementary. The total mass and
weight of the stationary magnetic system with the outer radius equal to that of the prototype but
optimized for NdFeB magnets would be 5.2 kg and 51 N respectively. If the system is scaled by a
factor of k, the mass and the weight apparently will be scaled by the factor of k 3 . Figures D.5 and
D.6 show achievable ratios of the load capacity and stiffness vs. the bearing weight with current
sizes and with the sizes scaled up by a factor of two. Figures D.7 and D.8 show similar ratios
calculated with respect to the rotor weight only.

Stiffness/Rotor weight, 1/mm

APPENDIX D. LOADING CHARACTERISTICS

127

50
40
30
20

Current sizes
Current sizes x2

10
0

100

200

300

400

Rotational frequency, Hz

Load Capacity/Rotor weight

Figure D.7: Predictions of the stiffness vs. rotor weight ratios achievable with the sizes of the
current prototype and when the sizes are scaled up by a factor of two.

140
120
100
80
60

Current sizes
Current sizes x2

40
20
0

100

200

300

400

Rotational frequency, Hz

Figure D.8: Predictions of the load capacity vs. rotor weight ratios achievable with the sizes of the
current prototype and when the sizes are scaled up by a factor of two.

Appendix E

Experimental details
E.1

Measurement of the rotating coil inductances and resistances

The estimation of the rotating coil inductances and resistances were complicated by low magnitudes of their values and non-linearities caused by the presence of the iron cores in the additional
inductors. In this section, we describe how we obtained the values which we used for the bearing
parameter estimation.
The resistance R was measured by passing a 2 A dc current, IDC , through the assembly including a coil and an additional inductor and measuring the DC voltage drop, UDC , across the assembly.
The resistance was calculated as
UDC
.
R=
IDC
The inductance was obtained by feeding a 1 A rms 60 Hz ac current IAC into the assembly and
measuring the rms volt drop VAC across the assembly. The voltage and current (rms values) were
measured with digital multimeters. The inductance was calculated then as


1
UAC 2
L=
R2 .
2 60 Hz
IAC
Note, that a constant inductance model gives only an approximate description of the assembly
electromagnetic properties because of the significant non-linearities introduced by the iron core.
The effects of these non-linearities can be seen clearly on the oscillogram shown in Figure E.1.
These oscillogram was obtained by feeding a 1 A 60 Hz harmonic ac current, IAC , into the assembly and monitoring the voltage, VAC , across the assembly. The harmonic current was obtained by
connecting the assembly in series with a resistor which had much higher resistance than the assembly impedance and applying a harmonic voltage to the system. Note, that the voltage VAC is far
from being harmonic.
The parameters of the four coil+inductor assembles used in the prototype are summarized in the
Table E.1.

E.2

Measurement of the radial position of the rotor.

The radial position of the rotor was measured using two optical sensors. The dependencies of the
voltages on the sensor outputs vs. the distance from the target mounted on the rotor (blackened
128

APPENDIX E. EXPERIMENTAL DETAILS

129

Current / 2
Voltage x 2

Voltage (V), Current (A)

0.5

-0.5

-1

-0.01

0.01

0.02

0.03

Time, sec
Figure E.1: AC voltage and current oscillograms measured in the assembly comprising a coil and
an additional inductor.

Coil #
1
2
3
4

R ()
0.0538
0.0543
0.0540
0.0543

L (mH)
0.483
0.466
0.486
0.488

Table E.1: Measured parameters of the rotating coils used in the bearing prototype.

APPENDIX E. EXPERIMENTAL DETAILS

130

Output voltage, V

4.5
4
3.5
3
2.5
2
1.5
1
0

Displacement, mm

Figure E.2: Sensor output voltage vs. distance dependence measured in the x-direction.
cylinder) to the sensitive elements are shown in Figures E.2 and E.3.
It can be seen that the sensors were unsensitive to the displacements until some thresholds were
reached. It was noticed that the voltage/displacement dependencies after the thresholds could be

well approximated by the curves V (u) = a + bu + c u. Thus, the following fitted curves were
obtained for the X and Y position sensors respectively:

(E.1)
Vx (x) = 5.458 1.776x + 8.506 x 0.8 < x < 5.8;

Vy (y) = 13.784 2.597y + 13.66 y

2.3 < y < 6.3.

(E.2)

These curves are plotted against the measured voltage-vs.-displacement curves in Figures E.4 and
E.5.
Solving the equations (E.1) and (E.2) about x and y, we obtain the following equations for the
x and y as functions of Vx and Vy :

(E.3)
x = 8.396 0.563Vx 2.697 8.39460 1.7760Vx ;
y = 8.526 0.385Vy 4.745


1.97748 0.47323Vy .

(E.4)

APPENDIX E. EXPERIMENTAL DETAILS

131

Output voltage, V

3.5
3
2.5
2
1.5
1
0

Displacement, mm

Figure E.3: Sensor output voltage vs. distance dependence measured in the y-direction.

Output voltage, V

4.5
4
3.5
3
2.5

Experiment
Fit

2
1.5
1
0

Displacement, mm

Figure E.4: Sensor output voltage vs. distance dependence measured in the x-direction and its

approximation with Vx (x) = 5.458 1.776x + 8.506 x.

APPENDIX E. EXPERIMENTAL DETAILS

132

Output voltage, V

4.5
Experiment
Fit

4
3.5
3
2.5
2
1.5
1
0

Displacement, mm

Figure E.5: Sensor output voltage vs. distance curve measured in the y-direction and its approxi
mation with Vy (y) = 13.784 2.597y + 13.66 y.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

133

Appendix F

Measurement of the radial loading


characteristics of the bearing prototype.
1320 RPM
Z
0
0.5
1.5
2.5
3.5
4.5
5.5
6.5
7.5
8.5
9.5
10.5
11.5
12.5
13.5
14.5
15.5

X
4.20
4.20
4.20
4.20
4.20
4.20
4.20
4.20
4.20
4.20
4.20
4.20
4.20
4.20
4.20
4.20
4.20

Y
3.79
3.78
3.77
3.72
3.70
3.64
3.60
3.55
3.50
3.46
3.39
3.33
3.26
3.18
3.10
3.03
2.93

Z
16.5
17.5
18.5
19.5
20.5
21.5
22.5
23.5
24.5
25.5

X
4.20
4.20
4.20
4.20
4.20
4.20
4.20
4.20
4.20
4.20

Y
2.83
2.77
2.69
2.59
2.43
2.35
2.22
2.08
2.00
1.80

Table F.1: Raw data for calculating the radial force-displacement characteristic at 1320 RPM.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

134

F, N

1.5

0.5

0
0

0.5

Y, mm

1.5

Figure F.1: Force-displacement characteristic in the Y-direction (the direction perpendicular to the
force) at 1320 RPM. The fitting function is F = 0.091 + 0.897y.
Parameter
K , N/mm
, degrees
K, N/mm
, degrees

Value
0.897
900
1.360
41.290

Table F.2: Summary of the radial suspension parameters measured at 1320 RPM (close to the minimal stable levitation speed).

The rotor displacement in the direction of the applied force (X-direction) was not detected at
this speed, which was close to the minimal levitation speed. Consequently, the angle between the
rotor displacement and the external force, , was 900 .

APPENDIX F. MEASURED LOADING CHARACTERISTICS

135

1600 RPM
Z
0
2.5
4.5
6.5
8.5
10.5
12.5
14.5
16.5
18.5
20.5
22.5
24.5

X
4.22
4.20
4.20
4.18
4.16
4.15
4.13
4.11
4.10
4.09
4.08
4.07
4.06

Y
3.81
3.75
3.68
3.59
3.50
3.37
3.24
3.10
2.95
2.75
2.56
2.37
2.15

Table F.3: Raw data for calculating the radial force-displacement characteristic at 1600 RPM.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

Experiment
Fit

1.5

F, N

136

0.5

0
0

0.1

X, mm

0.2

Figure F.2: Force-displacement characteristic in the X-direction (the direction of the external force)
at 1600 RPM. The fitting function is F = 0.01 + 6.455x.
Parameter
K , N/mm
, degrees
K, N/mm
, degrees

Value
0.925
81.350
1.478
38.240

Table F.4: Summary of the radial suspension parameters measured at 1600 RPM.

Experiment
Fit

F, N

1.5

0.5

0
0

0.5

Y, mm

1.5

Figure F.3: Force-displacement characteristic in the Y-direction (the direction perpendicular to the
force) at 1600 RPM. The fitting function is F = 0.087 + 0.934y.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

137

2
Experiment
Fit

F, N

1.5

0.5

0
0

0.5

R, mm

Figure F.4: Force vs total rotor displacement, R =


fitting function is F = 0.084 + 0.925r.

1.5

X 2 + Y 2 , characteristic at 1600 RPM. The

80
70
60

+, 0

50
40
30
20
10
0

0.5

R, mm

1.5

Figure F.5: The angle between displacement and force vs rotor displacements at 1600 RPM.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

138

1800 RPM
Z
0
2.5
4.5
6.5
8.5
10.5
12.5
14.5
16.5
18.5
20.5
22.5
24.5

X
4.20
4.20
4.18
4.16
4.13
4.11
4.08
4.05
4.03
4.01
3.99
3.97
3.95

Y
3.80
3.76
3.70
3.62
3.52
3.41
3.30
3.17
3.02
2.87
2.71
2.53
2.34

Table F.5: Raw data for calculating the radial force-displacement characteristic at 1800 RPM.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

Experiment
Fit

1.5

F, N

139

0.5

0
0

0.1

0.2

X, mm

0.3

0.4

Figure F.6: Force-displacement characteristic in the X-direction (the direction of the external force)
at 1800 RPM. The fitting function is F = 0.134 + 4.172x.
Parameter
K , N/mm
, degrees
K, N/mm
, degrees

Value
0.980
76.670
1.571
37.400

Table F.6: Summary of the radial suspension parameters measured at 1800 RPM.

Experiment
Fit

F, N

1.5

0.5

0
0

0.5

Y, mm

1.5

Figure F.7: Force-displacement characteristic in the Y-direction (the direction perpendicular to the
force) at 1800 RPM. The fitting function is F = 0.133 + 1.008y.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

Experiment
Fit

1.5

F, N

140

0.5

0
0

0.5

1.5

R, mm

Figure F.8: Force vs total rotor displacement, R =


fitting function is F = 0.113 + 0.980r.

X 2 + Y 2 , characteristic at 1800 RPM. The

80
70
60

+, 0

50
40
30
20
10
0

0.5

R, mm

1.5

Figure F.9: Angle between displacement and force vs rotor displacement at 1800 RPM.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

141

2000 RPM
Z
0
2.5
4.5
6.5
8.5
10.5
12.5
14.5
16.5
18.5
20.5
22.5
24.5

X
4.23
4.21
4.19
4.16
4.12
4.08
4.05
4.01
3.99
3.95
3.92
3.89
3.85

Y
3.82
3.78
3.72
3.64
3.55
3.44
3.33
3.20
3.08
2.94
2.80
2.62
2.45

Table F.7: Raw data for calculating the radial force-displacement characteristic at 2000 RPM.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

Experiment
Fit

1.5

F, N

142

0.5

0
0

0.1

0.2

0.3

X, mm

0.4

0.5

0.6

Figure F.10: Force-displacement characteristic in the X-direction (the direction of the external force)
at 2000 RPM. The fitting function is F = 0.06 + 3.031x.
Parameter
K , N/mm
, degrees
K, N/mm
, degrees

Value
0.982
70.810
1.634
34.590

Table F.8: Summary of the radial suspension parameters measured at 2000 RPM.

Experiment
Fit

F, N

1.5

0.5

0
0

0.5

Y, mm

1.5

Figure F.11: Force-displacement characteristic in the Y-direction (the direction perpendicular to the
force) at 2000 RPM. The fitting function is F = 0.099 + 1.038y.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

Experiment
Fit

1.5

F, N

143

0.5

0
0

0.5

1.5

R, mm

Figure F.12: Force vs total rotor displacement, R =


fitting function is F = 0.094 + 0.982r.

X 2 + Y 2 , characteristic at 2000 RPM. The

80
70
60

+, 0

50
40
30
20
10
0

0.5

R, mm

1.5

Figure F.13: Angle between displacement and force vs rotor displacement at 2000 RPM.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

144

2200 RPM
Z
0
2.5
4.5
6.5
8.5
10.5
12.5
14.5
16.5
18.5
20.5
22.5
24.5

X
4.24
4.21
4.18
4.15
4.10
4.07
4.03
3.98
3.94
3.90
3.85
3.81
3.77

Y
3.83
3.79
3.72
3.65
3.57
3.47
3.37
3.25
3.14
3.00
2.87
2.73
2.58

Table F.9: Raw data for calculating the radial force-displacement characteristic at 2200 RPM.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

Experiment
Fit

1.5

F, N

145

0.5

0
0

0.1

0.2

0.3

0.4

X, mm

0.5

0.6

0.7

Figure F.14: Force-displacement characteristic in the X-direction (the direction of the external force)
at 2200 RPM. The fitting function is F = 0.054 + 2.55x.
Parameter
K , N/mm
, degrees
K, N/mm
, degrees

Value
1.007
66.410
1.698
32.930

Table F.10: Summary of the radial suspension parameters measured at 2200 RPM.

Experiment
Fit

F, N

1.5

0.5

0
0

0.5

Y, mm

1.5

Figure F.15: Force-displacement characteristic in the Y-direction (the direction perpendicular to the
force) at 2200 RPM. The fitting function is F = 0.081 + 1.096y.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

Experiment
Fit

1.5

F, N

146

0.5

0
0

0.5

1.5

R, mm

Figure F.16: Force vs total rotor displacement, R =


fitting function is F = 0.076 + 1.007r.

X 2 + Y 2 , characteristic at 2200 RPM. The

80
70
60

+, 0

50
40
30
20
10
0

0.5

R, mm

1.5

Figure F.17: The angle between the rotor displacement and the external force, , at different rotor
displacements at 2200 RPM.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

147

2400 RPM
Z
0
2.5
4.5
6.5
8.5
10.5
12.5
14.5
16.5
18.5
20.5
22.5
24.5

X
4.24
4.22
4.19
4.14
4.10
4.05
4.00
3.95
3.90
3.85
3.80
3.75
3.69

Y
3.85
3.81
3.75
3.68
3.60
3.51
3.41
3.31
3.20
3.08
2.97
2.83
2.69

Table F.11: Raw data for calculating the radial force-displacement characteristic at 2400 RPM.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

Experiment
Fit

1.5

F, N

148

0.5

0
0

0.2

0.4

X, mm

0.6

0.8

Figure F.18: Force-displacement characteristic in the X-direction (the direction of the external force)
at 2400 RPM. The fitting function is F = 0.067 + 2.222x.
Parameter
K , N/mm
, degrees
K, N/mm
, degrees

Value
1.016
62.620
1.741
31.200

Table F.12: Summary of the radial suspension parameters measured at 2400 RPM.

F, N

1.5

Experiment
Fit

0.5

0
0

0.2 0.4 0.6 0.8

Y, mm

1.2 1.4 1.6

Figure F.19: Force-displacement characteristic in the Y-direction (the direction perpendicular to the
force) at 2400 RPM. The fitting function is F = 0.08 + 1.142y.

APPENDIX F. MEASURED LOADING CHARACTERISTICS

Experiment
Fit

1.5

F, N

149

0.5

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8

R, mm

Figure F.20: Force vs total rotor displacement, R =


fitting function is F = 0.077 + 1.016r.

X 2 + Y 2 , characteristic at 2400 RPM. The

80
70
60

+, 0

50
40
30
20
10
0

0.5

R, mm

1.5

Figure F.21: The angle between the rotor displacement and the external force, , at different rotor
displacements at 2400 RPM.

Appendix G

The high rotational speed assumption.


All the results presented in this dissertation are derived subject to the high rotational speed assumption first introduced in Section 3.2. Simplistically, this assumption implies that the rotation of the
rotor about its axis Z is much faster than lateral motions of the axis Z. It can be also said that the
high-speed assumption implies that we can neglect the transient time needed for the currents in the
rotating conductor to settle to a harmonic form during lateral motions of the rotational axis Z. More
details are given in Section 3.2.
Because of the high-speed assumption, the resulting theory is capable of reasonably accurate
prediction of the take-off (minimal stable levitation) speed only if the assumption is satisfied in the
proximity of this speed. This is normally the case in systems where the axial stabilization is passive,
such as shown in Figure 1.2, because the system becomes stable only if the rotational speed is high
enough for the stabilizing electromagnetic radial stiffness to overcome the destabilizing stiffness
introduced by the passive axial suspension. However, if an active system were used to control the
axial direction, then there might be no destabilizing radial stiffness. Moreover, the axial suspension
in this case can be designed to introduce stabilizing radial stiffness. This will significantly reduce the
take-off speed and might invalidate the high speed assumption, leading to inaccuracy in prediction
of the take-off speed.
In this dissertation we do not derive analytical stability conditions for systems which clearly do
not satisfy the high-speed assumption. However, in what follows, we illustrate through a simple
example how one can approach the analysis of such systems, especially if obtaining analytical
stability conditions turns out to be difficult. This example also gives more insight into the nature of
the high-speed assumption and gives some tools to evaluate how well the high-speed assumption is
satisfied.
We show that without making the high-speed assumption the open loop transfer function of
the proposed bearing can be presented as a product of two transfer functions. The first transfer
function is exactly the same as it was with the high-speed assumption, and it describes the low
speed dynamics of the lateral motions of the Z axis. The other transfer function reflects the highspeed dynamics of the system caused by the transient process of settling currents in the rotating
conductors to their harmonic shapes. The high-speed assumption implies that the characteristic
frequencies of these two transfer functions are located far apart from each other on the frequency
axis, in which case the influence of the second transfer function on the system stability is negligible.

150

APPENDIX G. THE HIGH ROTATIONAL SPEED ASSUMPTION.

G.1

151

An example of approaching the problem without making the high


rotational speed assumption.

The equations of the system dynamics can be obtained without making the high rotational speed
assumption. We illustrate it through a simple example, assuming that the rotating conductor G is
represented by a set of n separate coils. For simplicity, first we neglect the mutual inductances
between the coils. Later we indicate how they can be taken into consideration.
As before, we represent each vector {x; y} in the XY plane by a complex number x + iy. We
designate such complex vector representations with hats. In the rotor XR YR coordinate frame, we
choose the XR axis so that it passes through the midline of one of the coils. We designate this coil
as no. 1, and number the remaining coils sequentially in the direction of rotation. The Amperes
force produced by the k-th coil then can be presented in the XR YR coordinate frame as
k = Fk n
k ;
F

n
k = ei2

k1
n

(G.1)

Let the rotor be displaced from equilibrium by a vector r0 , the polar coordinate representation
for which in the Newtonian X0 Y0 coordinate frame is r0 = {r0 ;  }. The complex plane
 r0 = r0 ei .
representation of this vector is
The flux linked with the k-th coil, k , is:


2
(k 1) = (r0 mk ),
(G.2)
k = r0 cos t
n
where is a parameter that denotes the change in flux linkage with respect to position, and

 

cos  2
(k 1) + t
n
.
(G.3)
mk =
sin 2
n (k 1) + t
It is easy to see that in complex form,
k eit .
m
k = n

(G.4)

*
)
* )
 r0 n
 r0 n
 r0 (
nk eit ) =
k eit +
k eit ,

k =
2

(G.5)

Thus,

where means the complex conjugate.


The current in the k-th coil, Ik is given by the following equation:
L

dk
dIk
+ RIk =
.
dt
dt

(G.6)

Also, the magnitude of the force acting on the coil is proportional to the current in the coil with the
proportionality coefficient :
Fk = Ik .

(G.7)

Then, we can rewrite (G.6) in terms of the force:


R
dk
L dFk
+ Fk =
.
dt

dt

(G.8)

APPENDIX G. THE HIGH ROTATIONAL SPEED ASSUMPTION.

152

Substituting (G.5) yields


*
L dFk
R
d )
 r0 n
+ Fk =
k eit +
k eit .
r0 n
dt

2 dt

(G.9)

To obtain an equation for the force in the complex form comprising the information about the force
direction (in the rotating XR YR coordinate frame) as well as the force magnitude, we multiply both
parts of (G.9) by n
k and note (G.1). This yields:
R
L dFk
d )  2 it  it *
+ Fk =
k e + r0 e
.
r0 n
dt

2 dt

(G.10)

Now we resolve the force Fk into the Newtonian coordinate frame X0 Y0 using the rotation by
an angle t:
Fk = Fk0 eit .

(G.11)

Caution needs to be exercised when taking the derivative


d )  it * dFk0 it
dFk
=
=
e
i Fk0 eit .
Fk0 e
dt
dt
dt

(G.12)

Substituting (G.11) and (G.12) into (G.10) and subsequently multiplying (G.10) throughout by eit
yields:
L dFk0
+
dt

R iL

d )  2 it  it * it
e .
k e + r0 e
Fk0 =
r0 n
2 dt

(G.13)

Summing over all n coils and designating the total force as


F0 =

n


Fk0 ,

k=1

we get
L dF0
+
dt

R iL

d
F0 =
2 dt


 r0 eit

n



n
2k

 r0 e
+ n

it

eit .

k=1

Now we use the trick which we used several times in Chapter 3 and show that
n


n
2k

k=1

n


ei n (k1) = 0,

if

n 3.

k=1

To repeat, we use the equation for the sum of geometric series:


n

k=1

ei n (k1) =

e4i 1
4

ei n 1

Note that e4i 1 = 0,

.
and n 3,

ei n 1 = 0.

(G.14)

APPENDIX G. THE HIGH ROTATIONAL SPEED ASSUMPTION.

153

Therefore,
n


n 3,

ei n (k1) = 0.

k=1

Using this result, multiplying (G.14) throughout by i, and carrying out the differentiation on the
right, we get
iL

dF0
n
 r0 + iv0 ),
+ (L + iR)F0 = 2 (
dt
2

(G.15)

d 
r0 .
where v0 = dt
Let us define

= arctan

K=
=

R
;
L

(G.16)

2
R2 + (L)2
L

R2 + (L)2

(G.17)
(G.18)

Using these variables, we rewrite (G.15) as


ei( 2 )

dF0
 r0 ei K v0 ei( 2 ) .
+ F0 = K
dt

(G.19)

This equation can be also presented in the Cartesian coordinate form:








K sin cos
cos sin
sin cos

x.
(G.20)
x
F0 + F0 = K
sin cos
cos sin
cos sin
It is easy to see that balancing the real and imaginary parts of (G.19) gives the same result as
balancing the vector components in (G.20).
If we compare the right-hand part of (G.20) with the force diagram shown on Figure 3.6, we
note that the first term on the right-hand side of equation (G.20) is our position-dependent force
Fr , and the second term is the velocity-dependent force Fv . The only thing which was neglected
when making the high rotational speed assumption was the term on the left. It is easy to see that
=

1
cos
< .

(G.21)

Considering that monotonically goes to zero as 1/, while K approaches some saturation
value when increases, there must be some rotational speed when the high-speed assumption is
well justified. In the next section, we gain a little more insight to when the high-speed assumption
is justified.
This example can be easily extended to allow for cross coupling between the coils. To do this,
 , which consist of the directional complex vectors n
we introduce a vector n
k . If we have n coils,
 will be n. Each component of this vector is, in fact, a vector itself
the dimension of the vector n

APPENDIX G. THE HIGH ROTATIONAL SPEED ASSUMPTION.

154

represented by a complex number. Similarly we introduce an n-dimensional vector of the magnetic


fluxes through the coils, . Then we rewrite (G.5) as
=

*
)
 r0 n
 eit +
 eit .
r0 n
2

(G.22)

We also define the n-dimensional vectors of the currents I, the force magnitudes F and the forces,
 . Then we rewrite (G.6), (G.7), (G.8), and (G.9) by replacing n
 , I and F
F
k , Ik and Fk with n
respectively, and by replacing scalars L and R with n n matrices [L] and [R]. Thus, instead of
(G.9) we obtain
[L]

*
2 d ) 
dF
 r0 n
 eit +
 eit .
+ [R]F =
r0 n
dt
2 dt

(G.23)

The [L] and [R] matrices do not have to be diagonal.


We consider a particular case when all the coils are identical and located periodically. Because
of the system periodicity,
T ,
 T [L] = Leqv n
n

(G.24)

T ,
 T [R] = Reqv n
n

(G.25)

and

where Leqv and Reqv are some equivalent values of the coil inductance and the resistance. We show
this for the [L] matrix (the analysis for the [R] matrix is identical).
By definition:
 T = {e0 , . . . ei2
n

k1
n

, . . . ei2

n1
n

}.

 T [L] is
The k-th component of the row vector n
)

 T [L]
n

n


Ljk ei2

j1
n

= ei2

k1
n

j=1

n


Ljk ei2

jk
n

k
=n

j=1

n


Ljk ei2

jk
n

(G.26)

j=1

We note that Ljk depends only on the difference j k, or the angle 2 jk


n between the j and k
coils. We acknowledge this by writing


jk
Ljk = L 2
.
n
Then we note that the sum
n


i2 jk
n

Ljk e

j=1

n

j=1

jk
L 2
n

i2 jk
n

n1

j=0



j
j
ei2 n = Leqv
L 2
n

in (G.26) is independent of K. Substituting (G.27) into (G.26) yields


*
) *
)
T ,
 T [L] = Leqv n
n
k

(G.27)

APPENDIX G. THE HIGH ROTATIONAL SPEED ASSUMPTION.

155

which implies (G.24).


 T we get
Multiplying both sides of (G.23) by n
*
2 d )  T it 
d(
nT F )
 n
 e + r0 n
T n
 eit .
Leqv
+ Reqv (
nT F ) =
r0 n
dt
2 dt

(G.28)

0 in the rotor coordinate frame is


We note that the net force acting on the rotor, F
0 = n
T F .
F
We have already shown that
=
T n
n

n


ei n (k1) = 0.

k=1

Finally, it is easy to see that


 = n.
T n
n
Thus, we rewrite (G.28) as
Leqv

*
)
2
0
dF
 r0 eit .
0 = d n
+ Reqv F
dt
2 dt

(G.29)

The rest of the derivation is similar to that carried out earlier for independent coils. The only
difference is that the result will include equivalent values of the inductance and resistance, Leqv and
Reqv given by (G.24) and (G.25), instead of L and R.

G.2

Bode plot analysis.

G.2.1

Open loop model in the frequency domain.

To investigate the effects of on the system stability we consider Bode plots for the open loop
system. Let F(s) be the transfer matrix from x to F0 :
F0 = Fx.

(G.30)

x + C x,
and we write
In the presence of an external damping, F0 = m
x=

ms2

1
F.
+ Cs

(G.31)

Then, the open loop transfer function constructed from x to x, X , is


X =

ms2

1
11
1
F=
F,
+ Cs
C s d s + 1

where
d =

m
C

is the damping time constant.

(G.32)

APPENDIX G. THE HIGH ROTATIONAL SPEED ASSUMPTION.

156

From equation (G.20) we find


 

1 



s sin cos
sin cos
cos sin
F = K s
+I
+
cos sin
sin cos
cos sin
After some algebraic manipulation, we obtain


K
s2 + sin s + cos
s cos s + sin
.
F =
( 2 s2 + 2 sin s + 1) [ s cos s + sin ] s2 + sin s + cos
Lets consider the matrix


s cos s + sin
s2 + sin s + cos
.
F1 =
[ s cos s + sin ] s2 + sin s + cos
This is a square non-singular matrix. Therefore, it can be diagonalized by a similarity transformation. Considering that the matrix structure resembles that of a rotation matrix, it is easy to find a
transformation matrix R, the columns of which are the matrix eigenvectors:


1
i i
.
R=
1 1
2
Substituting s = i
and applying the transformation R yields


( +
)(
+ cos + i sin )
0
R F1 R =
,
0
(
)(
+ cos i sin )
where means complex conjugated transpose. (We replaced s with i
here because the result
cannot be reduced to any standard transfer function, for which Bode plots are tabulated. Note that
we use symbol
to distinguish the frequency as a variable from the rotational speed . )
Thus, we have decomposed our system into two, now independent, subsystems, each of which
can be analyzed independently. The transfer functions of the subsystems (multiplied by 1 since
the Bode plot analysis normally implies negative feedback) are
X1 =

)
1

*
1
(cos i sin )
K1
1+
(cos + i sin ) 2 2
C s d s + 1

s + 2 sin s + 1

(G.33)

X2 =

)
1

*
1 +
(cos + i sin )
K1
1
(cos i sin ) 2 2
C s d s + 1

s + 2 sin s + 1

(G.34)

We split these transfer functions into two parts:


X1 = X10 X1 ,

(G.35)

X2 = X20 X2 ,

(G.36)

and

where
X10 =

)
1
K1

* i
1+
e ;
C s d s + 1

(G.37)

APPENDIX G. THE HIGH ROTATIONAL SPEED ASSUMPTION.

157

X20 =

)
1

* i
K1
1
e ;
C s d s + 1

(G.38)

X1 =

1
(cos i sin )
;
2 s2 + 2 sin s + 1

(G.39)

1 +
(cos + i sin )
.
(G.40)
2 s2 + 2 sin s + 1
X10 and X20 are the same transfer functions as we would obtain neglecting , while X1 and
X2 are the corrections introduced when taking into account. (It is easy to see that X1 = 1 and
X2 = 1 if = 0.)
One can note that the differences between the transfer functions X10 , X20 , and the transfer
function of a mass-spring system with stiffness K, mass m and a damping coefficient C are the
 i
i
. It can be also noticed, that in spite of these multipliers, the
multipliers (1 + 
)e and (1 )e
gains of the transfer functions X10 and X20 go down with increase of
. Thus, similar to the massspring system, the stability of the proposed bearing when is not taken into account is determined
by the phase margin at the gain cross-over frequency,
c . We assume that the gain cross-over
frequency
c is much lower than the rotational frequency :
c << . (This is essentially our
high-speed assumption more explanations on how this relates to our ability to ignore will be

given later.) With this assumption, in the proximity of
c , the coefficients 1 + 
and 1 become
close to unity and the gains of the transfer functions X10 and X20 become equal to the gain of the
equivalent mass-spring system:
X2 =

|X20 | = |X10 | =

K
1

C 2 + m2
2

(G.41)

The important differences between the transfer functions X10 , X20 and that of the mass-spring
system are the phase shifts: + for X10 and for X20 . The second phase shift clearly reduces the
phase stability margin by the angle : the system stability is, therefore, defined by the phase margin
of X20 . The system becomes identical to the mass-spring system only if = 0.
If our assumption
c << is valid, then the crossover frequency of our system is close to that
of the mass-spring system and it can be found by making (G.41) equal to 1:
K
1

= 1.

c C 2 + m2
c2
The solution is

  
 2 0.5
1  C 2
K
1 C 4
+
+

c =
2 m
4 m
m
The standard triangle inequality implies that
  
 
 2
1 C 4
K
1 C 2 K
+

+ .
4 m
m
2 m
m

(G.42)

(G.43)

APPENDIX G. THE HIGH ROTATIONAL SPEED ASSUMPTION.


Consequently,
.

158

K
.
m


Thus, the condition K/m < is asufficient condition for
c << , which, however, tends to
be too conservative. Note that
0 = K/m is the natural frequency of the mass-spring system.
In the presence of strong damping, when


C
m

2


>> 2

K
m


,

the cross-over frequency can be approximated as

c
02 d = K/C.
Note that, with increasing C, the cross-over frequency goes to zero and the condition
c << is
likely to be satisfied.

G.2.2

Effects of on the system stability.

The time constant influences the system through the transfer functions X1 and X2 . Introduction of these two transfer matrices can destabilize the system in two ways: by reducing the phase
margin at the crossover frequency
c and by bringing the gain back above unity after the crossover
frequency
c is passed and when the phase is below 1800 . We will refer to these destabilizing
effects as phase margin effect and gain margin effect. These are considered below.
Phase margin effect.
Since we assumed that
c << and 1/ = / cos > , then
c << 1/ . Therefore, it is
reasonable to assume that the effects of X1 and X2 on the crossover frequency and on the phase
margin are small. This is another interpretation of our high rotational speed assumption. Because
there is a smaller phase margin in the transfer function X20 than in X10 , it is X2 which might ruin
the system stability, and not the X1 .
We introduce a dimensionless
=
/,

(G.44)

which indicates how far the frequency of the interest is from the rotational speed . As a trial point
we will often consider = 0.1, implying that the frequency of interest is one decade lower than the
rotational frequency of the rotor. It is easy to see that

= cos .

(G.45)

First we look at how X2 affects


c . The gain of X2 is

|X2 | =

1 + 2
2 + 2
cos
(1 2
2 )2 + 4 sin2 2
2

0.5
.

APPENDIX G. THE HIGH ROTATIONAL SPEED ASSUMPTION.

159

After some algebra it can be presented as


0.5 

0.5
1
1
|X2 | =
=
2 cos2 2 cos2 + 1
cos2 (1 )2 + sin2
Clearly, when < 1, |X2 | > 1, and it reaches its maximum at = 0. For example, if = 0.1,
|X2 |=0 = 1.11.
A small addition of gain to (G.41) due to X2 causes a small increase of
c and, consequently,
a small decrease of the phase margin.
Next we consider the direct change of phase due to X2 . Here we treat the nominator of X2
and the denominator separately. The phase of the nominator is



sin

(G.46)
n = arctan
;
0 < n < .
1 +
cos
2
The phase of the denominator is


2
sin
d = arctan
;
1 2
2

0 < d <

.
2

(G.47)

To establish the sign of the phase = n d introduced by X2 we consider the difference


tan n tan d =

sin
( 2
2 + 2
cos + 1)
< 0.
(1 2
2 )(1 +
cos )

The negative sign implies that < 0. Thus, introduction of leads to a reduction of the stability
phase margin and, consequently, to an increase of the take-off speed or damping needed for stability
at a given speed.
To get some quantitative estimate of the effect, we express in terms of (see equation (G.45)):




2 sin cos
sin cos
arctan
,
= arctan
1 + cos2
1 2 cos2
or, at small :
sin cos = 0.5 sin(2).
It is easy to see that
| | 0.5.

(G.48)

For example, at = 0.1


| | < 30 .

(G.49)

Gain margin effect.


The major problem one can expect with the gain stability margin is due to the second order oscillatortype transfer function in the denominators of X1 and X2 . If 0 then sin 0 and this transfer
function can be expected to have a pronounced peak at
= 1. In what follows, we show that this
is not going to be a problem.

APPENDIX G. THE HIGH ROTATIONAL SPEED ASSUMPTION.

160


Note that when
is close to 1/ , we cannot neglect the terms (1 + 
) and (1 ) in X10 and
X20 anymore (see (G.37) and (G.38)) . For the high speed analysis we withdraw these terms from
X10 and X20 and consider the functions
)
)

* 1
(cos i sin )
X1 = 1 +
;
(G.50)
X1 = 1 +

2 s2 + 2 sin s + 1

and
)
)

* 1 +
(cos + i sin )
X2 = 1
X1 = 1
.

2 s2 + 2 sin s + 1

(G.51)

Note that the remnants of the functions X10 and X20 are still (accurate to the phase) the transfer
functions of the open loop mass-spring system with the gain rolling off at high frequencies as
1/
2 .
We can already anticipate that the resonance at
= 1 is not going to be a problem. Indeed,
while at this frequency we have a significant gain increase because of the denominator, especially

when 0, we also have |1
(cos i sin )| 0 in (G.50), and |1 
| = |1 cos | 0
in (G.51). To show this more accurately, we substitute
/ = (see G.44) and
= cos (see
G.45) into (G.50) and (G.51). After some algebra, we get
|X1 |2 =

(1 + )2
,
+ 2 cos2 + 1

(G.52)

(1 )2
.
2 cos2 2 cos2 + 1

(G.53)

2 cos2

and
|X2 |2 =

We can easily find an upper bound on |X1 |:


0.5 
- 1+ (1 + )2
- |1 + |.
= -|X1 |
2 cos4 + 2 cos2 + 1
1 + cos2 -

(G.54)

The |X2 | can be presented as


|1 |
.
|X2 | = 
2
cos (1 )2 + sin2

(G.55)

Note that at = 1 (
= ) equation (G.55) gives us zero gain at synchronous speed for any
practical > 0. If = 0, |X2 | = 1 everywhere. If = /2, |X2 | = |1 |. At any intermediate
value of , |X2 | increases with from zero at = 1 at a rate not higher than (sin )1 |1 | until it
reaches the saturation value of (cos )1 .
Importantly, at any , |X2 | can be bounded as
|X2 | ||, if > 1;
|X2 | = 0, if = 1;
|X2 | |2 |, if < 1.

(G.56)

To show this for > 1 first, we form a difference


2 |X2 |2 =

(2 1) + cos2 3 ( 2)
.
(1 )2 cos2 + sin2

(G.57)

APPENDIX G. THE HIGH ROTATIONAL SPEED ASSUMPTION.

161

The sign of this difference is determined by the sign of the nominator, and may be negative only
if < 2, because the term cos2 3 ( 2) in the nominator in this case is negative. The absolute
value of this term increases with cos and cannot exceed 3 (2 ). Direct evaluation of (G.57)
with cos = 1 and 1 . . . 2 shows that this difference is never negative.
The proof for < 1 follows simply from the symmetry of (G.55) about the axis = 1.
Considering bounds (G.54) and (G.56), one should not expect problems with the gain margins
in the proximity of if it is well separated from the
c .
In conclusion, the high-speed assumption can be formulated as follows: The crossover frequency
c of an equivalent mass-spring system with the mass equal to the rotor mass m, damping
equal to the radial damping in the system C and stiffness equal to the in-plane stiffness K is much
lower than the rotational speed of the rotor . The
c is given by (G.43). As a rule of thumb, the
crossover frequency should be lower than the rotational speed by an order of magnitude or more.
In this case the direct change of the stability phase margin does not exceed 30 (see (G.49)). Direct
calculation of the system eigenvalues can be recommended to validate a design based on the high
rotational speed assumption.
As an example, in the prototype described in the dissertation, we have m=3.2 kg, C 143 Ns/m,
and K 1400 N/m as measured at the rotational speed =140 rad/s, which is slightly above the
take-off speed. Using (G.43), we calculate
c =9.6 rad/s, which is 15 times lower than . Thus,
according to (G.48), the direct change of the stability phase margin does not exceed 20 .
Thus, we conclude that the effect of is to increase the actual take-off speed relative to that
predicted using the high-speed assumption.

Das könnte Ihnen auch gefallen