You are on page 1of 229

A Thermoelastohydrodynamic Model of The Morton

Effect Operating in Overhung Rotors Supported by Plain


or Tilting Pad Journal Bearings
by

Avinash C. Balbahadur

Dissertation submitted to the Faculty of the


Virginia Polytechnic Institute and State University
in partial fulfillment of the requirements for the degree of
Doctor of Philosophy
in
Mechanical Engineering

APPROVED:

________________________
Dr. R.G. Kirk (Chair)

________________________ ________________________
Dr. M.E.F. Kasarda Dr. R.L. West

________________________ ________________________
Dr. B. Vick Dr. R.K. Kapania

February 27, 2001


Blacksburg, Virginia
Abstract

Unlike most instabilities, which are non-synchronous in nature, the Morton Effect

is a synchronous phenomenon. This thermal instability occurs primarily in overhung

rotors that are supported by fluid film bearings and is caused by differential viscous

shearing within the bearing lubricant. The Morton Effect has also gained much attention

within the last decade.

Prior studies of the Morton Effect have used complex analysis in the frequency

domain to model this instability. However, such an approach makes it difficult to develop

a user-friendly design tool for engineers. The current research employs a steady-state

analysis to predict the onset of the Morton Effect, and it uses an instability criterion which

is based on a threshold unbalance caused by a force equal to 15% of the weight of the

rotor. It is hoped that this method will provide a more easily adaptable platform for

design and analytical purposes.

The current model has demonstrated good agreement with other theoretical

models and experimental data. This agreement applies to rotors that are supported by

either plain or tilting pad journal bearings and it was found that a worse case scenario for

the Morton Effect would involve centered, circular and large-amplitude bearing orbits.

A test rotor was also designed and built. Initial experimental data revealed an

unusual instability that might have been caused by the Morton Effect.
Dedicated to My Parents:

Pooran and Savitri Balbahadur

iii
Acknowledgments

I am profoundly grateful for the advice provided by my advisor Dr. Kirk. Without

his guidance, this project would never have been accomplished. My gratitude also extends

to the members of my committee: Dr. Kasarda, Dr. Vick, Dr. West and Dr. Kapania for

their time and advice.

I am deeply indebted to Dr. Swanson for his help with the experimental system.

Also, many thanks to the technicians in the Mechanical Engineering Machine Shop. I am

especially grateful to Mr. James Dowdy for his precision work on various parts of the test

rotor. To Mr. Ben Poe and Mr. Jamie Archual goes my gratitude for rescuing me from a

quagmire of computer difficulties. I am also grateful for the invaluable assistance

provided by Ms. Eloise McCoy, Ms. Cathy Hill, Ms. Jackie Buhrdorf, Mr. Mike Harness

and the other members of the Mechanical Engineering Staff.

I would also like to thank my friends and relatives for all of their support. This

dissertation is dedicated to my parents, Mr. Pooran Balbahadur and Mrs. Savitri

Balbahadur, for their love and understanding throughout my academic endeavors.

iv
TABLE OF CONTENTS
ACKNOWLEDGMENTS.................................................................................................................... IV

TABLE OF CONTENTS .......................................................................................................................V

LIST OF FIGURES.............................................................................................................................VII

LIST OF TABLES ............................................................................................................................... IX

NOMENCLATURE ...............................................................................................................................X

1. INTRODUCTION ...............................................................................................................................1
1.1. LITERATURE REVIEW .......................................................................................................................2
1.2. MOTIVATION FOR RESEARCH .........................................................................................................10
1.3. RESEARCH OBJECTIVES ..................................................................................................................11
1.4. DEVELOPMENT OF THE CURRENT MODEL .......................................................................................11
2. THE MORTON EFFECT IN ROTORS WITH PLAIN JOURNAL BEARINGS..........................16
2.1. INITIAL MECHANICAL UNBALANCE.................................................................................................16
2.2. FILM THICKNESS FOR PLAIN JOURNAL BEARINGS ............................................................................17
2.3. STATIC EQUILIBRIUM JOURNAL POSITION .......................................................................................19
2.4. SYNCHRONOUS ORBIT ...................................................................................................................20
2.5. TEMPERATURE DISTRIBUTION ........................................................................................................24
2.6. THERMAL UNBALANCE ..................................................................................................................29
2.7. INSTABILITY CRITERION ................................................................................................................33
2.8. COMPUTER PROGRAM FLOWCHART ................................................................................................34
2.9. CASE STUDIES ...............................................................................................................................36
2.9.1. Keogh and Morton ...............................................................................................................36
2.9.2. Faulkner, Strong and Kirk...................................................................................................43
2.10. SOLUTION SENSITIVITY TO THE THRESHOLD UNBALANCE FORCE ..................................................51
3. THE MORTON EFFECT IN ROTORS WITH TILTING PAD JOURNAL BEARINGS ............53
3.1. FILM THICKNESS FOR TILTING PAD JOURNAL BEARINGS ..................................................................54
3.2. STATIC EQUILIBRIUM JOURNAL POSITION .......................................................................................64
3.2.1. Reynolds’ Equation...............................................................................................................65
3.2.2. Pressure Distribution ............................................................................................................67
3.2.3. Forces and Moments.............................................................................................................72
3.3. SYNCHRONOUS ORBIT ...................................................................................................................78
3.4. TEMPERATURE DISTRIBUTION ........................................................................................................82
3.5. CASE STUDIES ...............................................................................................................................88
3.5.1. de Jongh and Morton............................................................................................................88
3.5.2. de Jongh and van der Hoeven...............................................................................................97
3.5.3. Overhung Compressor ........................................................................................................103
4. DESIGN OF TEST ROTOR...........................................................................................................109
4.1. DESIGN PARAMETERS ..................................................................................................................110
4.2. ANALYSIS OF TEST ROTOR DESIGNS .............................................................................................114
4.3. TEST ROTOR COMPONENTS ..........................................................................................................120
4.3.1. Test Shaft............................................................................................................................121
4.3.2. Bearings..............................................................................................................................122
4.3.3. Coupling .............................................................................................................................123

v
4.3.4. Overhang Disk ....................................................................................................................124
4.3.5. Inboard Disk .......................................................................................................................126
4.4. ANALYSIS OF FINAL TEST ROTOR DESIGN .....................................................................................127
5. EXPERIMENTAL RESULTS FROM TEST ROTOR..................................................................140
5.1. ROTOR BALANCING .....................................................................................................................140
5.2. FIXED SPEED DATA......................................................................................................................142
5.3. ACCEL-DECEL DATA ...................................................................................................................145
6. CONCLUSIONS..............................................................................................................................160
6.1. CONCLUSIONS BASED ON THE THEORETICAL WORK.......................................................................160
6.2. CONCLUSIONS BASED ON THE EXPERIMENTAL RESEARCH ..............................................................161
7. SUGGESTED FUTURE RESEARCH ...........................................................................................162
7.1. SUGGESTED FUTURE THEORETICAL WORK....................................................................................162
7.2. SUGGESTED FUTURE EXPERIMENTAL RESEARCH ...........................................................................163
APPENDIX A: MATLAB GUI PROGRAM FOR PLAIN JOURNAL BEARINGS........................164

APPENDIX B: FORTRAN PROGRAM FOR TILTING PAD JOURNAL BEARINGS.................172

APPENDIX C: MATLAB GRAPHICS PROGRAM FOR FORTRAN TPJB PROGRAM DATA .207

REFERENCES....................................................................................................................................213

VITA....................................................................................................................................................216

vi
List of Figures
Figure 1. 1 - An Example of the Newkirk Effect........................................................................................2
Figure 1. 2 - An Example of the Morton Effect..........................................................................................6
Figure 1. 3 - Schematic Diagram of The Morton Effect Model used by Other Researchers.......................11
Figure 1. 4 - Schematic Diagram of Current Morton Effect Model ..........................................................13

Figure 2. 1 - Schematic Cross-Section Through a Plain Journal Bearing .................................................17


Figure 2. 2 - Schematic Diagram of the Synchronous Orbit .....................................................................22
Figure 2. 3 - Energy Flow in a Plain Journal Bearing ..............................................................................25
Figure 2. 4 - Thermally-induced Bend in a Rotor with an Overhung Mass...............................................30
Figure 2. 5 - Mechanical, Thermal and Resultant Unbalances .................................................................33
Figure 2. 6 - Flowchart for Plain and Tilting Pad Journal Bearing Programs ...........................................35
Figure 2. 7 - VT-FAST Model of the Keogh and Morton Symmetric Rotor.............................................36
Figure 2. 8 - Growth Factor Plot for Rotor (Keogh and Morton, 1994) ....................................................38
Figure 2. 9 - Unbalance Curves from Current Analysis of Keogh and Morton Rotor................................39
Figure 2. 10 - Forced Response Plots for Keogh and Morton Rotor (Unb. = 1 oz in) ................................40
Figure 2. 11 - Synchronous Orbit of the Keogh and Morton Rotor at 10505 RPM....................................41
Figure 2. 12 - Synchronous Orbit of the Keogh and Morton Rotor at 5730 RPM......................................42
Figure 2. 13 - VT-FAST Model of the Turbocharger studied by Faulkner et al. ......................................43
Figure 2. 14 - Unbalance Curves from Current Analysis of Turbine End of the Turbocharger studied by
Faulkner et al. .................................................................................................................................45
Figure 2. 15 - Synchronous Orbit from Turbocharger (Turbine End) at 9000 RPM..................................46
Figure 2. 16 - Synchronous Orbit from Turbocharger (Turbine End) at 11000 RPM................................47
Figure 2. 17 - Forced 9000 RPM Turbine End Orbit: The Ellipse is Centered..........................................48
Figure 2. 18 - Forced 9000 RPM Turbine End Orbit: The Centered Ellipse is Changed to a Circle with
Radius = Semi-Major Axis of Ellipse...............................................................................................49
Figure 2. 19 - Forced 9000 RPM Turbine End Orbit: The Centered Circular Orbit is Enlarged by a Factor
of Ten..............................................................................................................................................50

Figure 3. 1 - Schematic Diagrams of Tilting Pad Journal Bearings Showing Load-on-Pad and Load-
between-Pads Configurations...........................................................................................................54
Figure 3. 2- Schematic Diagram of a Cylindrical Pivot TPJB ..................................................................55
Figure 3. 3 - Triangles from Diagram of The Cylindrical Pivot TPJB ......................................................57
Figure 3. 4 - Comparison of Film Thickness Expressions for a 5-Pad Cylindrical Pivot TPJB..................63
Figure 3. 5 - Coordinate System for Reynolds’ Equation..........................................................................65
Figure 3. 6 - Pad Surface Showing Pressure Boundary Conditions...........................................................67
Figure 3. 7 - A 4-Node Isoparametric Quadrilateral Element...................................................................70
Figure 3. 8 - Comparison between Calculated and Actual Dimensionless Stiffness and Damping
Coefficients from Someya’s 5 -Pad Bearing .....................................................................................81
Figure 3. 9 - Temperature Distribution for Knight’s Bearing...................................................................86
Figure 3. 10 - Film Thickness, Pressure and Temperature at the Centerline of Knight’s Bearing (Table 3.
6).....................................................................................................................................................87
Figure 3. 11 - VT-FAST Model of deJongh and Morton Centrifugal Compressor Rotor...........................89
Figure 3. 12 - Critical Speed Maps for VT-FAST Model and Actual Compressor ....................................91
Figure 3. 13 - Unbalance Curves for the Non-Driven End of the deJongh and Morton Compressor (Before
Modifications) .................................................................................................................................92
Figure 3. 14 - Unbalance Curves for the Driven End of the de Jongh and Morton Compressor (Before
Modifications) .................................................................................................................................93
Figure 3. 15 - Unbalance Curves for the Non-Driven End of the de Jongh and Morton Compressor (After
Modifications) .................................................................................................................................95
Figure 3. 16 - Unbalance Curves for the Driven End of the de Jongh and Morton Compressor (After
Modifications) .................................................................................................................................96

vii
Figure 3. 17 - VT-FAST Model of a Compressor Analyzed by de Jongh and van der Hoeven ..................97
Figure 3. 18 - Critical Speed Maps for VT-FAST Model and an Actual Compressor Analyzed by de Jongh
and van der Hoeven .......................................................................................................................100
Figure 3. 19 - Unbalance Curves for the Non-Driven End of a Compressor Analyzed by de Jongh and van
der Hoeven (Before Modifications) ................................................................................................101
Figure 3. 20 - Unbalance Curves for the Non-Driven End of a Compressor (with Heat Sleeve Barrier)
Analyzed by de Jongh and van der Hoeven ....................................................................................102
Figure 3. 21 - Synchronous Vibration Data obtained from an X-Probe at the NDE of the Overhung
Compressor ...................................................................................................................................104
Figure 3. 22 - Synchronous Vibration Data obtained from an X-Probe at the DE of the Overhung
Compressor ...................................................................................................................................105
Figure 3. 23 - Unbalance Curves for the NDE of the Overhung Compressor..........................................106
Figure 3. 24 - Unbalance Curves for the DE of the Overhung Compressor.............................................107

Figure 4. 1- Schematic Diagram of Existing Test Rig............................................................................111


Figure 4. 2- Schematic Diagram Showing the Modified Rotor Configuration for The Morton Effect .....112
Figure 4. 3 - Average Morton Effect Instability Speeds and Critical Speeds for the Designs ..................116
Figure 4. 4 - |Angle| between Mechanical and Thermal Unbalances for the Different Designs ...............117
Figure 4. 5 - Static Eccentricity Ratio at Nmort for Each of the Designs...................................................118
Figure 4. 6 - Orbit Size at Nmort for the Various Designs........................................................................119
Figure 4. 7 - 303 Stainless Steel Shaft ...................................................................................................121
Figure 4. 8 - Bearing Housing Components...........................................................................................123
Figure 4. 9 - Schematic Diagram of the Kop-Flex Coupling Showing the Required Modifications.........124
Figure 4. 10 - Overhang Disk Design ....................................................................................................125
Figure 4. 11 - Inboard Disk Design .......................................................................................................126
Figure 4. 12 - The Final Design for the Morton Effect Test Rotor..........................................................127
Figure 4. 13 - Picture of the Actual Morton Effect Test Rotor................................................................128
Figure 4. 14 - Picture of the Morton Effect Rotor and Associated Systems .............................................129
Figure 4. 15 - VT-FAST Model of The Morton Effect Rotor..................................................................130
Figure 4. 16 - Predicted 1st Damped Critical Speed for the Morton Effect Rotor ....................................132
Figure 4. 17 - Predicted 2nd Damped Critical Speed for the Morton Effect Rotor....................................133
Figure 4. 18 - Predicted 3rd Damped Critical Speed for the Morton Effect Rotor ....................................134
Figure 4. 19 - Predicted 4th Damped Critical Speed for the Morton Effect Rotor ....................................135
Figure 4. 20 - Response at Probes 3,4 due to a 1oz in. Unbalance at the NDE Overhang Mass C.G. ......136
Figure 4. 21 - Unbalance Curves for the Non-Driven End of the Morton Effect Test Rotor (Final Design)137
Figure 4. 22 - Unbalance Curves for the Driven End of the Morton Effect Test Rotor (Final Design).....138

Figure 5. 1 - Vibration Readings from Probe 5 (Before and After Balancing) ........................................141
Figure 5. 2 - Speed and Probe 3 Response for a Fixed Speed Data Plot ..................................................144
Figure 5. 3 - Probe 3 Response Data for 5000-9000 RPM Acceleration-Deceleration (Oil Inlet
Temperature = 100 oF)...................................................................................................................145
Figure 5. 4 - Probe 5 Response Data for 5000-9000 RPM Acceleration-Deceleration (Oil Inlet
Temperature = 100 oF)...................................................................................................................146
Figure 5. 5 - Cascade Plot for Vibrations Shown in Figure 5. 3 .............................................................147
Figure 5. 6 - Cascade Plot for Vibrations Shown in Figure 5. 4 .............................................................148
Figure 5. 7 - Shaft Centerline Position (from Probes 3 and 4) near NDE Bearing .................................149
Figure 5. 8 - Shaft Centerline Position (from Probes 5 and 6) near DE Bearing .....................................150
Figure 5. 9 - Probe 3 Response Data for 5000-9000 RPM Acceleration-Deceleration (Oil Inlet
Temperature = 78.3 oF)..................................................................................................................152
Figure 5. 10 - Probe 5 Response Data for 5000-9000 RPM Acceleration-Deceleration (Oil Inlet
Temperature = 78.3 oF)..................................................................................................................153
Figure 5. 11 - Probe 3 Response Data for 5000-9500 RPM Acceleration-Deceleration (Oil Inlet
Temperature = 100 oF)...................................................................................................................154

viii
Figure 5. 12 - Probe 5 Response Data for 5000-9500 RPM Acceleration-Deceleration (Oil Inlet
Temperature = 100 oF)...................................................................................................................155
Figure 5. 13 - Synchronous Response at the NDE Bearing of de Jongh and Morton’s LP/IP Compressor
(de Jongh and Morton, 1994).........................................................................................................157

List of Tables
Table 1. 1 - Comparison between the Newkirk Effect and the Morton Effect .............................................7
Table 1. 2 - Comparison between Other and Current Morton Effect Models ............................................13

Table 2. 1 - Data for Keogh and Morton rotor .........................................................................................37


Table 2. 2 - Data for Turbine End of Faulkner, Strong and Kirk Turbocharger........................................44
Table 2. 3 - Predicted and Actual Instability Speed Values (RPM) for Different Threshold Unbalance
Force Levels ....................................................................................................................................51

Table 3. 1- Input Data for Comparison of Film Thickness Expressions....................................................62


Table 3. 2 - Basis Function Parameters....................................................................................................71
Table 3. 3 - Input Data for Bearing Number 10 (5-Pad TPJB), Test Number 10 (Someya, 1989) .............75
Table 3. 4 -Computed ε0 and ψa0 as a Function of Mesh Size for Someya’s Bearing ................................76
Table 3. 5 - Input Data for a 5-Pad LBP TPJB (Knight, 1990).................................................................77
Table 3. 6 - Input Data for a 5-pad LBP TPJB (Knight, 1990) .................................................................84
Table 3. 7 - Data Generated from Bearing Described in Table 3. 6 ..........................................................85
Table 3. 8 - Input Data for deJongh and Morton Compressor Rotor Before Modifications........................90
Table 3. 9 - Input Data for de Jongh and Morton Compressor Rotor After Modifications.........................94
Table 3. 10 - NDE Data for Compressors Analyzed by de Jongh and van der Hoeven..............................99
Table 3. 11 - Data for Overhung Compressor ........................................................................................103

Table 4. 1- Description of Rotor Designs...............................................................................................115


Table 4. 2 - Summary of Design Analysis..............................................................................................120
Table 4. 3 - Data for Morton Effect Rotor..............................................................................................131

ix
Nomenclature
Dimensions: M = Mass, L = length, t = time, T = temperature

Ax, Ay amplitude parameters for elliptic orbit (L)

cl lubricant specific heat capacity (L2 t-2 T-1)

Cb radial bearing clearance (L)

Cp radial pad clearance (L)

e eccentricity (L)

E Young’s Modulus (M L-1 t-2)

E& b rate of energy transfer to bearing (M L-2 t-3)

E& j rate of energy transfer to journal (M L-2 t-3)

E& lub rate of energy storage in lubricant (M L-2 t-3)

E& visc rate of visc. energy dissipation (M L-2 t-3)

fp fractional angular position of pad pivot

h film thickness (L)

H heat transfer coefficient (M t-3 T-1)

I area moment of inertia (L4)

L bearing length (L)

Ld disk overhang length (L)

m preload factor

md mass of overhung disk (M)

Npads number of pads

x
Ob bearing center

Oj journal center

Op pad center of curvature

P hydrodynamic pressure within the bearing (M

L-1 t-2)

P0 lubricant supply pressure (M L-1 t-2)

Rb bearing radius (L)

Rj journal radius (L)

Rp inner radius of curvature of pad (L)

Re Reynolds’ number

Rec critical Reynolds’ number (Re > Rec =>

turbulent)

tp Pad thickness

T journal circumferential temperature (T)

T0 lubricant supply temperature (T)

Tamb ambient temperature (T)

uj surface velocity of journal (L t-1)

U resultant unbalance (M L)

Um mechanical unbalance (M L)

Ut thermal unbalance (M L)

Uthr threshold unbalance (M L)

W rotor weight (M L t-2)

xi
Wb bearing load (M L t-2)

x,y,z local coord. system with origin on journal

surface

X,Y,Z fixed coord. System with origin at bearing

center

yd thermal deflection of disk (L)

z axial dimension (L)

a thermal conductivity of the journal (T-1)

β thermoviscosity coefficient (T-1)

δ pad tilt angle

∆p angular dimension of pad

ε eccentricity ratio

θ circumferential angle

θc angle for line-of-centers

θj angle to journal center

θp pivot angle for pad

µ lubricant viscosity (M L-1 t-1)

µ0 lubricant supply viscosity (M L-1 t-1)

ξ angle used in PJB film thickness expression

ρl lubricant density (M L-3)

τ lubricant shear stress (M L-1 t-2)

φ angle between thermal and mechanical

xii
unbalances

φx, φy phase parameters for elliptical orbit

y thermal bend angle

ya attitude angle

ω angular journal speed (t-1)

13
1. Introduction

The term ‘rotordynamics’ or ‘rotor dynamics’ consists of two words: ‘rotor’ and

‘dynamics’. ‘Rotor’ refers to objects which spin or rotate while ‘dynamics’ deals with the

motion of objects subjected to imposed forces. Hence, ‘rotordynamics’ can be defined as

the study of spinning objects which are subjected to imposed forces.

Since the invention of the wheel in Mesopotamia circa 3500-3000 B.C., humans

have been fascinated by spinning objects. However, it was not until 1869 that Rankine

wrote the first article on rotordynamics (Kirk, 1998). In this article Rankine concluded

that operation above the critical speed would be unstable. This hypothesis was later

disproved by DeLaval (in 1895) when he demonstrated that a steam turbine was capable

of running above the first critical speed. A theoretical explanation of this phenomenon

was offered by the English dynamicist, H. H. Jeffcott in 1919. The Jeffcott rotor model

consisted of a mass-less elastic shaft on rigid supports with an unbalanced disk at the mid-

span of the shaft. This model showed that the rotor response peaked only at the critical

speed and attenuated to stable values at speeds above and below this critical. Even though

this model is simplistic, and only accounts for a single critical speed, it is still used today in

order to obtain a conceptual understanding of basic rotordynamics.

1
1.1. Literature Review
Just seven years after Jeffcott developed his rotor model, Newkirk published a

paper (Newkirk, 1926) on a thermally-induced problem which was observed in the rotor

of a water-wheel generator. Newkirk concluded that the cause of the problem was a

stator element which rubbed against the shaft to produce a hot spot (Figure 1. 1).

hot spot caused by


mechanical rotor-stator contact
unbalance
*
shaft
disk

bearing bearing

(a) Mechanical unbalance in phase with hot spot (unstable)

hot spot caused by


rotor-stator contact

disk
shaft

* mechanical
unbalance

bearing bearing

(b) Mechanical unbalance out of phase with hot spot (stable)

Figure 1. 1 - An Example of the Newkirk Effect

2
During this operation, the rotor is executing synchronous whirl, i.e. the shaft or journal is

rotating at the same speed as the rate at which the shaft center is orbiting the bearing

center. This configuration ensures that a particular area of the shaft surface rubs against

the stator piece to create a hot spot, while the diametrically opposite portion of the shaft

never experiences this frictional contact and remains at the ambient temperature. Such a

temperature difference causes a thermal gradient to develop across the shaft and

eventually leads to thermal bending. Newkirk observed that if the shaft was running

below its first critical speed then the mechanical unbalance enhances the thermal bend

caused by the hot spot, as shown in Figure 1. 1(a). This leads to increased shaft vibrations

which promote more contact with the stator. As a result, the system becomes unstable via

a positive feedback mechanism. On the other hand, if the shaft is running above the first

critical speed (Figure 1. 1(b)), then the mechanical unbalance is roughly out of phase with

the hot spot and inhibits the unbalance produced by the thermal bend vector. Hence, the

system shown in Figure 1. 1(b) will be stable which is concordance with Newkirk’s

observations. It should be noted that the Newkirk Effect is not limited to occurring below

the first critical speed. Most of the later researchers, who are referenced in the next

paragraphs, have witnessed this problem in high speed turbomachinery.

R.P. Kroon and W.A. Williams later gave a mixed qualitative and quantitative

analysis (Kroon and Williams, 1939) of the Newkirk Effect, but it was decades before a

purely quantitative model (Dimarogonas, 1973) was established. Dimarogonas computed

the static thermal bow due to an arbitrary heat input and then combined this bow with the

system dynamics to yield a dynamic thermal bow. The final model consisted of two

nonlinear differential equations which had to be solved numerically. The requirement of a

3
digital computer to solve this problem may be one reason why it took about 30 years

before quantitative results were obtained. The solution published by Dimarogonas

indicated that the Newkirk Effect consisted of 3 modes: a spiralling mode in which

vibration increased in time, an oscillating mode which consisted of constant-amplitude

sinusoidal vibrations, and a constant mode with a fixed time-independent vibration level.

Dimarogonas concluded that the main factor which determines the vibration mode is the

phase difference between the static mechanical bow and the dynamic thermal bow relative

to a coordinate system fixed to the rotor.

The Newkirk Effect was then studied by Kellenberger (Kellenberger, 1980) who

found this phenomenon affecting gas-cooled turbogenerators. Frictional interaction

between the turbogenerator rotor and stator, via a seal ring, produced the hot spot

required for thermal bending. Unlike Dimarogonas, Kellenberger obtained linear

equations by making some simplifying assumptions, e.g. the shaft thermal bow is linearly

proportional to the shaft temperature. He also found that the interaction between the

thermal and mechanical unbalances was an essential driving force behind the Newkirk

Effect. In addition, he discovered that the ratio of the heat flow into the shaft to the heat

flow out of the shaft determined the system stability, i.e. if this ratio was above some

threshold curve then the system would yield unstable spiral vibrations. Schmied (Schmied,

1987) extended the work of Dimarogonas and Kellenberger to include multi-bearing

rotors by using a finite element code. He found that the hot spots could also be caused by

slipring brushes rubbing against the rotor shaft. More recent work concerning the

Newkirk Effect has been done by Goldman and Muzynska (Goldman and Muzynska,

1995) and Goldman, Muzynska and Bently (Goldman et al., 1998). These researchers

4
have noted that the speed of the transient thermal effects is very slow compared to the

rotor vibration velocities. As a result, the thermal problem can be uncoupled from the

rotor vibrations. They have also accounted for inelastic impacts during rotor-to-stator

contact.

Schmied (Schmied, 1987) indicated that unstable spiral vibrations could also arise

from hot spots developing within the fluid-film bearing. Research done by Keogh and

Morton (Keogh and Morton, 1994) confirms the existence of such instabilities which

occur primarily in overhung rotors (Figure 1. 2). These unstable vibrations constitute the

phenomenon currently referred to as the Morton Effect.

5
hot spot caused by
mechanical
uneven viscous shearing
unbalance
within bearing

*
inboard rotor mass shaft

overhung disk

bearing bearing

(a) Mechanical unbalance in phase with hot spot (stable)

hot spot caused by


uneven viscous shearing
within bearing

mechanical
inboard rotor mass shaft unbalance

*
overhung disk

bearing bearing

(b) Mechanical unbalance out of phase with hot spot (unstable)

Figure 1. 2 - An Example of the Morton Effect

The Morton Effect occurs when the journal is executing a synchronous orbit around the

bearing center. This orbit causes one portion of the journal surface to always be at the

minimum film thickness, while a diametrically opposite section of the journal surface is

always at the maximum film thickness. Lower film thickness areas are generally

associated with higher viscous shear stresses which produce higher temperatures. As a

result, a hot spot will develop on the journal surface exposed to the minimum film

6
thickness region and a cold spot will be formed on the surface at maximum film thickness.

This temperature difference leads to a temperature gradient developing across the journal.

If the hot spot is 180 degrees out of phase with the overhung unbalance (Figure 1. 2(b)),

and if the temperature gradient and the magnitude of the unbalance is adequate, thermal

bending will occur. Under these conditions the bent shaft will decrease the bearing

clearance and elevate the thermal gradient. The increased temperature gradient will then

initiate more thermal bending. These actions describe a positive feedback mechanism

which will drive the system unstable.

Both the Newkirk Effect and the Morton Effect involve the development of a

temperature gradient across the journal which eventually leads to unstable thermal bending

via a positive feedback mechanism. Some of the other similarities and differences between

these two phenonmena are summarized in Table 1. 1.

Table 1. 1 - Comparison between the Newkirk Effect and the Morton Effect

Parameter Newkirk Effect Morton Effect

Rotor Motion synchronous whirl synchronous whirl

Overhung Configuration not required required

Mechanical Unbalance in phase with hot spot out of phase with hot spot

Cause of Hot Spot rotor-to-stator rub differential viscous shearing

Location of Hot Spot outside bearing within bearing

The first thorough analysis (Keogh and Morton, 1994) of the Morton Effect

considered a symmetric overhung rotor supported by plain journal bearings. A time-

7
dependent thermal bend was first calculated by combining the heat transfer equations with

the dynamic equations of the rotor. This thermal bend was then transformed to the

frequency domain where it was incorporated into a positive feedback loop. The stability

characteristics of this loop were then obtained by plotting Nyquist and eigenvalue graphs

as a function of rotor speed. Keogh and Morton thus obtained a range of instability

speeds for which the real part of the eigenvalues were positive.

A more practical insight into the Morton Effect was obtained by deJongh and

Morton (deJongh and Morton, 1994) who investigated a synchronous vibration instability

encountered in a centrifugal compressor which was used in an off-shore gas-lift

application. This compressor was mounted on two oil-lubricated, tilting pad journal

bearings and had a mass of about 450 kg (990 lbm). DeJongh and Morton found that the

compressor rotor exhibited rapid and continuously-growing vibrations around 11400

RPM. They also noticed that when the rotor speed was reduced the unstable vibrations

gradually subsided. This hysteresis in response indicated that the problem could not have

been caused by running too close to a critical speed. In any case, the nearest critical was

around 14500 RPM which was sufficiently separated from the operating speed. Such

behavior implied that the instability could be thermal in nature, so the researchers removed

the mechanical and labyrinth seals and restarted the rotor. Unfortunately, the unstable

vibrations persisted which implied that rotor-to-stator contact (via the seal rings) could

not have been responsible for this anomaly. Hence, the Newkirk Effect could be ruled out

as the potential culprit. The compressor was finally stabilized by reducing the overhang

unbalance which was achieved by replacing some of the steel parts with lighter ones made

of titanium. This occurrence of a non-Newkirk thermal instability, which was sensitive to

8
overhang unbalance, seemed to imply that the Morton Effect could be the source of the

problem. DeJongh and Morton confirmed this hypothesis by building a test rotor with

identical dynamic characteristics as the actual rotor. They measured the temperature

difference across the journal and showed that a rise in this transverse temperature gradient

coincided with the onset of the unstable vibrations.

Faulkner, Strong and Kirk (Faulkner et al., 1997) also encountered the Morton

Effect operating in a large turbocharger. The synchronous orbit and the overhung turbine

wheel caused a hot spot to develop within the turbine-end bearing. This problem was

solved by undercutting the 3-axigroove bearing to increase the bearing operating

eccentricity. Such a configuration forced a less-centered orbit which caused the hot spot

to be alternately heated and then cooled. As a result, the average temperature of the hot

spot diminished and the thermal gradient was reduced. The thermal bending was therefore

decreased and the ensuing vibrations were attenuated to an acceptable level.

Another method for decreasing the thermal gradient from the Morton Effect was

given by de Jongh and van der Hoeven (de Jongh and van der Hoeven, 1998). These

researchers installed a heat barrier sleeve around the portion of the shaft within the

bearing. This sleeve had a much lower thermal conductivity than the shaft and partially

insulated the shaft from the heat generated within the bearing. Consequently, the

temperature difference across the shaft was reduced and the influence of the Morton

Effect was mitigated.

Recently, there has been great interest in the Morton Effect. Some researchers

(Gomiciaga and Keogh, 1999) have used CFD techniques to obtain a better estimate of

9
the journal temperature distribution. Others, such as Larsson (Larsson I and II, 1999),

have extended the work of Keogh and Morton to include tilting pad journal bearings.

1.2. Motivation for Research


The Newkirk Effect has been around for more than 70 years and has been well-

studied and is better understood than the Morton Effect. In addition, the Morton Effect

has ignited an abundance of interest in the past few years. As a result, there is a greater

precedence for current research to be directed at the Morton Effect than at the Newkirk

Effect.

So far, most of the analytical approaches to the Morton Effect have been identical

and concentrate on a complex analysis in the frequency domain. Perhaps it is possible to

develop a simpler and more user-friendly model which can easily be adapted to become a

design tool for industrial applications. State-of-the-art research has also failed to make

adequate comparisons between theoretical models and experimental data. Hence, there is

room for research in this particular area. Furthermore, the relative novelty of the Morton

Effect means that there are still unanswered questions concerning its mechanism. For

instance, Why is the overhung configuration necessary? Does the orbit shape affect the

mechanism? These and other questions give further incentive for more research to be

concentrated on this particular phenomenon.

10
1.3. Research Objectives
The objectives of this research are as follows:

(1) Develop a simple model for the Morton Effect.

(2) Use the model to obtain a better understanding of the Morton Effect.

(3) Compare the model with other theoretical research and experimental data.

(4) Initiate experimental work on the Morton Effect.

1.4. Development of the Current Model


Most of the prior models of the Morton Effect have involved complex stability

analyses (Keogh and Morton, 1994; Larsson I and II, 1999), and the typical outline of

these models is shown in Figure 1. 3.

synchronous circumferential
Initial Bend temperature Thermal Bend
orbit
+ distribution
Dynamic Thermal
Σ
Fluid Film
Rotor Model Bearing Model Rotor Model
+

Figure 1. 3 - Schematic Diagram of The Morton Effect Model used by Other

Researchers

In these models, an initial deformed rotor configuration is assumed and is then used as an

input in a dynamic model of the rotor system. This dynamic model produces a

synchronous orbit, i.e. the rate at which the shaft rotates about its own center is equal to

the frequency at which the shaft center rotates about the bearing center. Such an orbit

causes a circumferential temperature distribution to develop on portions of the shaft within

11
the bearings. A thermal bend, which depends on the temperature distribution and the

rotor thermal characteristics, is subsequently initiated. This thermal bend affects the rotor

dynamics and gives rise to positive feedback behavior. The stability of the resulting

feedback transfer function can then be analyzed by using Nyquist plots or another

equivalent method.

The initial mechanical and thermal bends, which are used in the models by other

researchers, can be related to unbalances. An unbalance arises when some of the rotor

mass lies off the axis of rotation. As a result, a centrifugal force, which is equal to the

product of the unbalance (mass x distance b/w mass and rotation axis) and the square of

the angular speed, is created. This unbalance affects the response of the rotor system and

therefore determines the size of the synchronous orbit. For a perfectly balanced rotor

(unbalance = 0), there is no orbit and the shaft simply rotates at the static equilibrium

position. The main point of this discussion is that the bends in the previous models can be

replaced with unbalances.

In real machines satisfactory performance is established by using specified

vibration, or unbalance, threshold levels which are given in standards published by various

organizations. The American Petroleum Institute (API) is one such organization and, in

1979, it provided API standard 617 for centrifugal compressors (American Petroleum

Institute, 1979). This standard said that, after final balancing, the maximum allowable

residual unbalance should be equal to the unbalance caused by a force which is 10% of the

journal static load. In other words, the maximum allowable centrifugal force on the

compressor is 10% of the static load. Such a criterion determines whether the rotor will

be “stable” or “unstable” during operation, i.e. if the unbalance level exceeds the threshold

12
level given by the 10% criterion then the rotor will be “unstable” and will have to be re-

balanced. This notion of practical stability will be utilized in the current model.

The current model for the Morton Effect is shown in Figure 1. 4 and it first

requires an estimate for the initial mechanical unbalance in the rotor.

initial synchronous circumferential stable/


thermal
unbalance orbit temperature unstable
unbalance
Dynamic Fluid Film distribution Thermal Instability
Rotor Model Bearing Model Rotor Model Criterion

Figure 1. 4 - Schematic Diagram of Current Morton Effect Model

This unbalance is then put into the dynamic rotor model to obtain the synchronous orbit.

A fluid film bearing model is then used to establish the temperature distribution which

leads to the thermal unbalance. Finally, an unbalance threshold criterion is used to

determine the system stability.

Figure 1. 3 and Figure 1. 4 illustrate two different models for the Morton Effect

and these two representations can be compared as shown in Table 1. 2.

Table 1. 2 - Comparison between Other and Current Morton Effect Models

Parameter Other Models Current Model

Domain frequency time

Instability Criterion Nyquist criterion unbalance > threshold level

Model Type transient, feedback steady-state, no feedback

Number of Calculations several per rotor speed one per rotor speed

13
The most fundamental difference between the two models is that prior representations

transform all equations to the frequency domain, while the current model operates in the

time domain. Hence, the current model avoids the complications encountered when

applying the Laplace Transform to non-linear partial differential equations.

A time domain model also allows a simpler instability criterion -- the resultant

unbalance has to exceed some threshold value -- to be employed. However, other models

have to use the more complicated Nyquist stability theorem. In addition, the steady-state

nature and the fewer calculations of the current model make it less computationally

intensive than the other models. Previous models have to update the thermally-deformed

rotor shape after each successive iteration. As a result, these models tend to involve more

calculations than the current representation.

Before the current model is further developed, Figure 1. 4 must be examined in

greater detail. Once a method for estimating the initial mechanical unbalance has been

ascertained, the synchronous orbit must be established. This involves superimposing a

dynamic orbit on the static equilibrium position of the journal. Obtaining this equilibrium

position requires a knowledge of the lubricant film thickness which depends on the

geometry of the bearings supporting the rotor. The dynamic orbit is then calculated using

VT-FAST (Virginia Tech - Front-end Automated Simulation of Turbomachinery) which is

a rotordynamics software package that is currently used in both the classroom and

industry. This information can then be used in conjunction with an energy equation to

obtain the circumferential temperature distribution (within the bearings) around the shaft

or journal. Such a temperature distribution leads to a thermal deflection which causes a

thermal unbalance. The resultant of the thermal and the initial mechanical unbalances can

14
then be determined and compared with a threshold value. If the resultant unbalance is

greater than the threshold value then the system will be unstable and the Morton Effect

will occur in that rotor.

In summary, the current model involves the following steps:

1. Estimate the initial mechanical unbalance.

2. Determine the film thickness for the particular bearing.

3. Obtain the static equilibrium position of the journal.

4. Use VT-FAST to get the dynamic orbit. This orbit is then combined with the

equilibrium position to yield the total synchronous orbit.

5. Develop an energy equation to obtain the circumferential temperature distribution.

6. Calculate the thermal unbalance.

7. Compare the resultant unbalance with a pre-defined instability threshold criterion.

Like previous research, the current model will focus on rotors supported by plain journal

bearings (Keogh and Morton, 1994; Faulkner et al., 1997) and tilting pad journal bearings

(deJongh and Morton, 1994; de Jongh and van der Hoeven, 1998). The simpler plain

journal bearing case will be examined first.

15
2. The Morton Effect in Rotors with Plain
Journal Bearings

The development of the Morton Effect model for rotors supported by plain journal

bearings will follow the steps outlined in the introduction section.

2.1. Initial Mechanical Unbalance


Each rotor system has its own weight distribution, stiffness and initial shaft

deformation. As a result, the amount and the location of the mechanical unbalance is very

difficult to predict in an arbitrary rotor system. Therefore, a nominal initial mechanical

unbalance (Um) will be assumed. This mechanical unbalance will be defined as the

unbalance created from a centrifugal force equal to 10% of the total static rotor weight

(W). Um will occur when the rotor is running at its maximum continuous operating speed

(ωMCOS), and can be mathematically defined as:

01
.W
eq2. 1 Um =
ω 2MCOS

This unbalance will act at an angle of zero degrees with respect to a coordinate

system on the rotor, and will be located at the center-of-gravity of the overhung mass.

16
2.2. Film Thickness for Plain Journal Bearings
The film thickness expression for a plain journal bearing has been derived

extensively in the literature. The following derivation is based on work done by Cameron

(Cameron, 1966). Cameron considered a cross-section through a plain journal bearing

(Figure 2. 1) which showed the shaft or journal center (Oj) displaced from the bearing

center (Ob) by some eccentricity, e.

Lubricant

Line-of-centers B

Rb δ
h

Ob
Rj
e
ξ
Oj
Journal

ψa

Figure 2. 1 - Schematic Cross-Section Through a Plain Journal Bearing

17
Applying the sine rule to the triangle ObOjB shown in Figure 2. 1 gives:

R
sin(ξ)
e e
eq2. 2 = b ⇒ sin(δ) =
sin(δ) sin ξ Rb

where Rb = bearing radius and ξ = circumferential angle between the film thickness line

and the line-of-centers. The above equation can be used with a trigonometric identity to

yield:

e2
eq2. 3 cos(δ) = 1− sin 2 (δ) = 1− sin 2 (ξ)
Rb 2

This means that cos(δ) is approximately equal to 1 because e2/Rb2 << 1 (typical values are

e = 2 mils or 50.8 µm and Rb = 2 inches or 5.08 cm, which gives e2/Rb2 = 10-6. The

maximum value of sin2(θ) is unity). The approximation for cos(δ) can be used in

conjunction with Figure 2. 1 to obtain an expression for the film thickness, h. From the

triangle in Figure 2. 1:

eq2. 4 h = e cos(ξ) + R b cos(δ) − R j ≈ e cos(ξ) + R b − R j

Defining the bearing clearance as Cb = Rb-Rj and the eccentricity ratio ε = e/Cb enables the

film thickness expression to be reduced to the following equation:

eq2. 5 h = C b + e cos(ξ) = C b + εC b cos(ξ)

Therefore,

eq2. 6 [
h = C 1+ ε cos(ξ)
b ]

18
2.3. Static Equilibrium Journal Position
The film thickness expression (eq2. 6) can be used to obtain the static eccentricity

ratio (ε0 - eccentricity ratio before elliptical orbit is superimposed) and the attitude angle

(ψa0 - angle between vertical static bearing load and the line-of-centers) which are needed

to establish the static equilibrium journal position. These parameters can be derived by

solving a reduced form of the Reynolds equation for the lubricant pressure distribution,

and subsequently integrating this pressure distribution over the journal surface area within

the bearing. The resultant force must balance the bearing load, Wb, so that static

equilibrium can be maintained.

These derivations have been done in detail by Cameron (Cameron, 1966) who

arrived at the following equation involving the static eccentricity ratio:

Wb C 2b π 2 ε 20 ε 40
eq2. 7 = +
3 
µωR j L3  2
161− ε 0 
 24
1− ε 0 
   

Equation eq2. 7 is valid for a narrow plain journal bearing, which means that the bearing

length, L, is small compared to the journal diameter, 2Rj. Furthermore, the lubricant

viscosity, µ, depends on the lubricant temperature which is a function of the journal

angular speed, ω. Unfortunately, the value of the eccentricity ratio is needed to obtain this

temperature distribution. A simple solution to this predicament is to use an effective

viscosity which will be defined later.

An 8th order polynomial can be obtained from the simplification of eq2. 7 and then

solved for the eccentricity ratio, ε0. The solution is presented within the computer code in

19
the appendices. Cameron (Cameron, 1966) also gives an expression for the static attitude

angle of a narrow bearing:

2
π 1− ε 0
eq2. 8 tan ψ =
a0 4 ε
0

By using equations eq2. 7 and eq2. 8, the static equilibrium position of the journal in the

plain journal bearing can be ascertained.

2.4. Synchronous Orbit


During its synchronous orbit, the journal traverses an elliptical path about its static

equilibrium position, Oj0 (see Figure 2. 2). The rate of whirl must equal the rate of

rotation in order for the orbit to be described as synchronous. Once this configuration is

achieved, a general point, P, on the surface of the journal will also travel along an elliptical

path. If the orbit was non-synchronous then the path taken by P would be an elliptical

shape with superimposed lobes.

With reference to Figure 2. 2, the path taken by the general point (P(x,y)) on the

journal surface can be mathematically described as:

 
( )
(a) x = e cos θ j0  + A cos ωt + φ x + R cos(ωt + λ )
0 x j
eq2. 9
   
(b) y = e sin θ j0  + A sin ωt + φ y  + R sin(ωt + λ )
0 y j

The geometry in Figure 2. 2 can be related to that of Figure 2. 1 by using the following

relationship:

20
eq2. 10 ξ = π −  ωt + λ − θ j 

where,

    
 e 0 sin θ j0  + A y sin ωt + φ y  
eq2. 11 θ = tan − 1  
j  e cos θ  + A cos ωt + φ 
 0  j0  x ( )
x 

and


eq2. 12 θ = +ψ
j0 2 a0

21
Y

P(x,y)

Journal Rj
Oj ωt+λ Bearing Sleeve

Aysin(ωt+φy)

e
ξ
Oj0 γ
e0 Axcos(ωt+φx)
θj
θj0
Ob
X

Elliptic Orbit
Centered at Oj0

Figure 2. 2 - Schematic Diagram of the Synchronous Orbit

From the geometry in Figure 2. 2, the location of the hot and cold spots can be

determined. The hot spot (H) will be defined as the point closest to the bearing sleeve at

time t = 0. In this configuration, it is also assumed that γ = ωt+λ so that Oj0, Oj and P=H

all lie on the same straight line. Hence at time t = 0:

A
y sin φ y 
eq2. 13 γ = γ 0 = tan −1   =λ
 A x cos φ x 

22
The time-dependent locations of the hot spot (H(xH,YH)) and the diametrically-opposite

cold spot (C(xC,yC)) can now be obtained by using the definition of γ0:

(a ) x
H 0 x ( j ) (
= e cos θ j0  + A cos ωt + φ x + R cos ωt + γ 0 )
eq2. 14
(b) y = e sin θ j0  + A sin ωt + φ y  + R sin(ωt + γ 0 )
H 0 y j

(a ) x
C 0 x ( j ) (
= e cos θ j0  + A cos ωt + φ x + R cos ωt + γ 0 + π )
eq2. 15
(a ) y = e sin θ j0  + A sin ωt + φ y  + R sin(ωt + γ 0 + π )
C 0 y j

However, for plain journal bearings, calculations are usually done in the ξ-coordinate

system and it is necessary to use eq2. 10 to obtain the locations of the hot and cold spots

in this reference:

eq2. 16 ξ = π −  ωt + γ 0 − θ j 


H

eq2. 17 ξ = π −  ωt + γ 0 + π − θ j 


C

It should be noted that the parameters Ax, Ay, φx and φy can be obtained directly

from the forced response program in VT-FAST (Virginia Tech - Front-end Automated

Simulation of Turbomachinery). VT-FAST is a rotordynamics software package that is

currently used in both the classroom and industry. The transfer matrix option in VT-

FAST was chosen to execute the forced response program, and the required stiffness and

damping coefficients for plain journal bearings were obtained from the literature (Vance,

1988):

23
Wb  2   
(a ) K = 4 2π + 16 − π 2  ε 2 Q(ε)
xx Cb  
 2 2  2 4 
Wb π − π + 2π ε + 16 − π  ε Q( ε)
2
(b) K xy = −
Cb   1/ 2
ε1− ε 2 
 2  2 2  2 4
Wb π π +  32 + π  ε + 216 − π  ε Q( ε)
(c) K yx = −
Cb   1/ 2
ε1− ε 2 
 2  2 2  2 4 
Wb 4 π +  32 + π  ε + 216 − π  ε Q( ε )
(d ) K yy =
 
Cb 1− ε 2 
 

 2  1/ 2 π 2 + 2 π 2 − 8 ε 2 Q( ε )


Wb 2π1− ε     
(e) C xx =
C bω ε
Wb  2  2  2 
(f ) C xy = − 8 π + 2 π − 8 ε Q( ε)
C b ω  
(g) C yx = C xy
 2  2 2 2 4 
Wb 2π π + 2 24 − π  ε + π ε Q( ε)
(h) C yy =
Cbω   1/ 2
eq2. 18 ε1− ε 2 
1
where Q(ε ) =
 2  2  2  3/ 2
π + 16 − π  ε 

The complete synchronous dynamic orbit and the location of the hot and cold spots can

now be obtained.

2.5. Temperature Distribution


The next step in the model is to obtain the circumferential temperature distribution

at the journal surface within the bearing. In order to obtain this temperature distribution,

an energy equation must first be formulated and solved. The physical domain, over which

this energy equation is to be applied, can be represented as shown in Figure 2. 3.

24
E& b

Bearing

h E& lub x E& lub x + dx

E& visc
Lubricant

ωRj

E& j
Journal

Figure 2. 3 - Energy Flow in a Plain Journal Bearing

Figure 2. 3 represents a simple model for energy flow in a plain journal bearing. It

assumes negligible axial heat flow and steady-state conditions within the control box. It

should be noted that the control box encompasses a portion of the journal surface. Hence,

the temperature within the box will be the same as the circumferential journal surface

temperature. Another key assumption in this model is that the final temperature of the

journal is obtained by averaging the distributions from each of the dynamic positions. As a

result, the average journal temperature distribution can be obtained by considering several

points in the synchronous orbit.

25
Conservation of energy can generally be stated as:

Energy Generation Rate = Energy Outflux Rate + Energy Accumulation Rate

The steady-state assumption means that the accumulation rate is zero. Therefore,

applying conservation of energy to Figure 2. 3 yields:

 
eq2. 19 E& visc =  E& lub x + dx − E& lub x + E& j + E& b  + 0
 

Substituting the appropriate terms into the above equation gives:

ωR j  dT  ωR j
ωR τdxdz = ρ hdz c  T + dx − ρ hdz cT
eq2. 20 j l 2 l dx  l 2 l
( )
+ fHdxdz T − Tamb + (1− f ) Hdxdz T − Tamb ( )

The rate of viscous energy dissipation, E& visc , depends on the speed of the journal surface

and the viscous shear stress, τ, which acts over an area dxdz (z = axial dimension). This

energy source term heats up the advecting lubricant which has a density of ρl and a

specific heat capacity of cl. A fraction, f, of the remaining heat is transferred to the journal

while the rest is assumed to be lost to the bearing housing and surroundings. It is also

assumed that the total heat loss is equal to Hdxdz(T-Tamb), where H = heat transfer

coefficient and Tamb = average ambient journal temperature = average ambient bearing

temperature. After simplification the above energy equation reduces to:

ωR j
eq2. 21 ωR j τ = ρ l h
2
cl
dT
dx
(
+ H T − Tamb )

If the lubricant is Newtonian, then the shear stress will be given by τ = µdu/dy.

Furthermore, a linear velocity profile can be assumed, i.e. the shear strain rate du/dy =

26
ωRj/h. This Petroff type of simplification has been found to give reasonable accuracy

(Cameron, 1966). It should also be noted that since h << Rj (journal radius), x ~ Rjξ. The

net result of these substitutions yields the following equation:

 
dT 2H  2 HT 2µωR 2j 
eq2. 22 + T− amb +  =0
dξ ρ l c l ωh  ρ l c l ωh ρ c h 2 
 l l 

At this point, the temperature rise above the ambient temperature can be defined as
~
( )
T = T − Tamb . Once these substitutions are used, equation eq2. 22 can be rearranged to

give :

2
2 H ~ 2µωR j
~
dT
eq2. 23 + T− =0
dξ ρ l c l ωh ρl c l h 2

The viscosity in eq2. 23 is a function of temperature. This dependency makes it difficult

to obtain the viscosity, a priori, without knowledge of the temperature distribution. As a

result, an effort has been made to derive an effective lubricant viscosity (µ) that can be

used in all equations. The first step was to replace Cb with h, and substitute the known

lubricant supply viscosity (µ0) for µ, in equation eq2. 23. Hence, an approximate 1D

equation for the temperature profile can be developed:

~
dT ~
+ AT − B = 0

eq2. 24
2H 2µ 0 ωR 2j
A= B=
ρ l c l ωC b ρ l c l C 2b

27
With the boundary condition T( ξ= 0) = T = T − T
~ ~
(T0 = lubricant supply
0 0 amb

temperature), equation eq2. 24 can be solved to give:

B ~ B
eq2. 25 ()
T ξ = +  T0 −  e − Aξ
~
A  A

Now, an average temperature rise (∆Tav) can be derived by assuming that this temperature

rise occurs over one half of the bearing circumference:

 B ~ 
∆Tav = T( π) − T0 ≈ Tmax − T0 =  − T0  1− e − Aπ 
~ ~ 
eq2. 26
A   

The effective viscosity may be defined according to the exponential Reynolds formulation:

− β∆Tav
eq2. 27 µ = µ 0e

where µ0 is the lubricant viscosity at the supply temperature, T0, and β = thermoviscosity

coefficient. Some researchers have indicated that this expression is not extremely accurate

(Cameron, 1966; Williams, 1994). However, it is simple and provides sufficient accuracy

for use in computation (Pinkus, 1990; Monmousseau et al., 1997).

Equation eq2. 23 can now be solved numerically (see the Matlab program in the

appendices). The required boundary condition sets the temperature at maximum film

thickness (the lowest temperature) to the value of the lubricant supply temperature, i.e.

~
( ~
)
T h max  ≡ T ξ = 0 = T0 . This equation (eq2. 23) is solved at several different points in the
~

dynamic synchronous orbit, and the temperature difference between the hot and the cold

spots are extracted from each solution. These temperature differences are then averaged

to give an overall mean temperature difference (∆T) between the hot and cold spots. The

28
resultant ∆T will be used to compute the thermal unbalance generated by the Morton

Effect.

2.6. Thermal Unbalance


A thermally-induced bend in the overhung rotor can be represented as shown in

Figure 2. 4. The static bending moment (M) in the journal can be expressed as:


eq2. 28 M = EI
dz

where E = Young’s Modulus, I = area moment of inertia and y = bend angle as shown in

Figure 2. 4.

29
Ld

L/2 L/2

hot spot

Journal z 2Rj
y
cold spot

Overhung Mass
Bearing

hot spot

yb
Journal
yd1 yd
cold spot ψb

Bearing

Overhung Mass

Figure 2. 4 - Thermally-induced Bend in a Rotor with an Overhung Mass

30
The maximum stress (smax) in the shaft can be written as:

R jM
eq2. 29 σ max =
I

If this maximum stress is primarily caused by thermal effects, then:

eq2. 30 σ max = Eα∆T

where a = thermal conductivity of the journal and ∆T is the overall mean temperature

difference between the hot and the cold spots. It is assumed that ∆T is independent of

axial position within the bearing. Furthermore, the small deflections allow the thermal and

mechanical problems to be uncoupled.

Equation eq2. 28 can be substituted into eq2. 29, and eq2. 29 and eq2. 30 can be

equated to give:

dψ α∆T
eq2. 31 =
dz R j

For a given journal speed and orbit, the right hand side of equation eq2. 31 is essentially

constant. Therefore, eq2. 31 can be integrated between the start of the bearing -- where it

is assumed that the bend angle is zero -- and some arbitrary axial distance, z (see Figure 2.

4), to give:

dy α∆T
eq2. 32 ψ= = z
dz R j

where y = deflection of the journal.

This equation can be evaluated at the end of the bearing to give the bend angle at this

location:

31
α∆T
eq2. 33 ψb = L
Rj

Also, eq2. 32 can be integrated over the bearing to give the deflection (yb) at the end of

the bearing:

1 α∆T 2
eq2. 34 yb = L
2 Rj

The additional deflection (yd1) at the overhung mass location can then be calculated using

the small angle assumption:

 L  L
eq2. 35 y
d1  d 2  ( )
=  L −  sin ψ b ≈  L d −  ψ
 2 b

By substituting eq2. 33 into eq2. 35 and by adding up eq2. 34 and eq2. 35, the total

thermal deflection at the overhang can be obtained:

α∆T
eq2. 36 yd = LL d
Rj

Equation eq2. 36 gives us the thermal deflection of the overhung mass. This deflection

can be converted into a thermal unbalance, Ut:

eq2. 37 U t = md y d

where md = overhung mass.

32
2.7. Instability Criterion
The resultant unbalance (U) from Um and Ut can be represented as shown in Figure

2. 5.

Um

U Line joining
ωt cold spot
to hot spot
θCH
X
Ut

Figure 2. 5 - Mechanical, Thermal and Resultant Unbalances

With reference to Figure 2. 5, the mechanical and thermal unbalances can be added

vectorially to produce the resultant unbalance (U) which can be represented as follows:

eq2. 38 (
U = U 2t + U 2m − 2U t U m cos ωt − θ CH )
The angle (ωt-θCH) is actually calculated by averaging similar angles at several different

points in the synchronous orbit.

A threshold unbalance (Uthr) for this synchronous thermal instability has to be

defined. This unbalance was assumed to be caused by a force equal to 15% of the rotor

weight and it can be mathematically expressed as:

33
015
. W
eq2. 39 U =
thr
ω2

where W = rotor weight and ω = variable angular journal speed. The value of 15% was

chosen because it gave the best results with the cases studied (see section 2.10).

The rotor will be unstable whenever U exceeds Uthr,. As a result, the threshold

speed for instability (ωthr) occurs when U = Uthr. This instability criterion can be obtained

graphically from the intersection of the U vs. ω and the Uthr vs. ω curves, and it determines

the onset of the Morton Effect in the rotor system.

2.8. Computer Program Flowchart


The flowchart shown in Figure 2. 6 is applicable to both the plain and tilting pad

journal bearing cases. At the beginning of the program, the rotor speed is updated and the

effective viscosity is calculated. This information is then used to establish a static

equilibrium position. The mechanical unbalance (Um) is next used in VT-FAST to get the

dynamic orbit which is combined with the static position to obtain the total synchronous

orbit. The temperature difference between the hot spot and the cold spot is then

calculated at equal time intervals in the orbit. By averaging these temperature differences,

the overall mean temperature difference (∆T) between the hot spot and the cold spot is

acquired. Next, the unbalances (Ut, U and Uthr) are computed and the process is repeated

for each rotor speed. U and Uthr can then be compared over the speed range to determine

whether the Morton Effect will occur in the rotor.

It should be noted that two programs were written. A Matlab Program (Appendix

A) is to be used with rotors supported by plain journal bearings, while a Fortran program

34
(Appendix B) is to be used with rotors supported by tilting pad journal bearings. The

graphics for the Fortran program is handled by the Matlab code in Appendix C.

START

Speed

Effective Viscosity

Static Equilibrium Position

Synchronous Orbit (VT-FAST)

Thot-Tcold

NO Finished #
of avg. ?

YES
∆T, Unbalances

NO Finished #
of spd. ?

YES
Plot U, Uthr vs. Speed

END

Figure 2. 6 - Flowchart for Plain and Tilting Pad Journal Bearing Programs

35
2.9. Case Studies

2.9.1. Keogh and Morton


In their paper (Keogh and Morton, 1994), Keogh and Morton present a theoretical

analysis of a symmetric rotor which is shown in Figure 2. 7.

Figure 2. 7 - VT-FAST Model of the Keogh and Morton Symmetric Rotor

This rotor consists of a shaft with a step variation in diameter. The shaft is supported by

two identical plain journal bearings and a disk of constant mass is placed, at an overhung

position, on either shaft end. The pertinent data for this rotor is summarized in Table 2. 1.

36
Table 2. 1 - Data for Keogh and Morton rotor

Parameter SI Units EG Units


Lubricant Properties
Density ρl 850 kg/m3 0.031 lbm/in3
Specific heat capacity cl 2000 J/kg/oC 0.478 Btu/lbm/oF
o o
Supply temperature T0 45.0 C 113 F
Supply viscosity µ0 0.0095 Pa s 1.38 µreyn
β
o -1 o -1
Thermovisc. coeff. 0.029 C 0.016 F

Bearing Properties
Length L 0.035 m 1.38 in
Radial clearance Cb 1.00E-04 m 3.94 mils
Heat transfer coefficient H 50 W/m2/oC 2.43e-5 hp/in2/oF
Bearing load Wb 2500 N 562 lbf
Journal radius Rj 0.050 m 1.97 in
α
o -1 o -1
Journal C.T.E 1.10E-05 C 6.11E-06 F

Rotor Properties
Rotor weight W 5000 N 1124 lbf
Overhung mass md 32.3 kg 71.3 lbm
Overhang Distance Ld 0.16 m 6.4 in
Max. Cont. Op. Speed ωMCOS 1047 rad/s 10000 RPM
Initial Mech. Unbalance Um 4.54e-4 kg m 0.63 oz in

The Keogh and Morton model examined the stability of a complex rotor thermal bend

angle. From this complex analysis, they were able to calculate the eigenvalues of the rotor

system and determine the corresponding growth factors (σ = real part of eigenvalue).

Since the system response depends on est, a positive growth factor would indicate a

theoretical instability. The growth factor plot for the Keogh and Morton case is shown in

Figure 2. 8.

37
Figure 2. 8 - Growth Factor Plot for Rotor (Keogh and Morton, 1994)

From Figure 2. 8, Keogh and Morton concluded that the critical speed range, in which

rotor thermal bending can occur, is between 1023 rad/s (9769 RPM) and 1086 rad/s

(10371 RPM).

The current analysis used the data in Table 2. 1 to predict a synchronous thermal

instability range from 1047 rad/s (10001 RPM) to 1206 rad/s (11521 RPM) which

overlaps with the Keogh and Morton range. In addition, the current model indicates

another range of instability after 1286 rad/s (12277 RPM). These instability ranges were

obtained from the intersection of the resultant unbalance and threshold unbalance curves

shown in Figure 2. 9.

38
Figure 2. 9 - Unbalance Curves from Current Analysis of Keogh and Morton Rotor

The agreement of the current model with the Keogh and Morton data is not excellent, but

it is still acceptable. One reason for some discrepancy could be the difficulty encountered

in fitting the Keogh and Morton viscosity information with an exponential profile. In

some instances (particularly at higher speeds) the viscosity values from the curve fit were

greater than the given data. This disagreement could have contributed to higher

temperatures and larger thermal unbalances at the elevated speeds. As a result, the

predicted instability range would be more prolific than the actual one.

The driving force behind the instability hump in Figure 2. 9 is the high overall mean

temperature (∆T) between the hot and cold spots. As mentioned above, higher viscosity

values could have contributed to these large ∆T values. However, the rotor dynamics

39
plays the more important role in generating these high ∆Ts which lead to greater thermal

and resultant unbalances. Figure 2. 10 shows that the rotor has a critical speed (4th

critical) near 10500 RPM.

Figure 2. 10 - Forced Response Plots for Keogh and Morton Rotor (Unb. = 1 oz in)

This lateral critical speed creates large amplitude orbits in this region. One such orbit (at

10505 RPM) is illustrated in Figure 2. 11.

40
Figure 2. 11 - Synchronous Orbit of the Keogh and Morton Rotor at 10505 RPM

The lower left and the upper right corners of this orbit show that, in these regions, the

rotor is running very close to the radial bearing clearance, Cb. This configuration implies

that the film thickness at the hot spot (x) would be very low, while the diametrically-

opposite cold spot (o) would experience higher film thickness values. Since a lower film

thickness is associated with higher viscous dissipation and higher temperatures, a strong

thermal gradient will develop across the journal. Such a gradient will cause the ∆T value

to be high and promote a thermal instability.

On the other hand, lower amplitude orbits (see Figure 2. 12), which are further

away from the critical speed, keep the rotor away from the clearance circle and are less

likely to be associated with high ∆T values. Hence, the orbit shown in Figure 2. 12 would

41
not give rise to a thermal bending instability. The lower speed of the 5730 RPM orbit also

reduces the amount of viscous dissipation and this further stabilizes this orbit relative to

the 10505 RPM one.

Figure 2. 12 - Synchronous Orbit of the Keogh and Morton Rotor at 5730 RPM

42
2.9.2. Faulkner, Strong and Kirk
Faulkner, Strong and Kirk (Faulkner et al., 1997) did an experimental study of a

large turbocharger which is shown in Figure 2. 13.

Compressor
Turbine
Impeller
Disk

Figure 2. 13 - VT-FAST Model of the Turbocharger studied by Faulkner et al.

The turbocharger had a centrifugal compressor impeller at one end and a radial inflow

turbine disk at the other end. This machine was supported by two 3 axial-groove journal

bearings which consist of a plain journal bearing with 3 small grooves cut along the

bearing length. For simplicity, this type of bearing will be approximated as a plain journal

bearing. Other relevant information on the turbine end of this turbocharger is shown in

Table 2. 2.

43
Table 2. 2 - Data for Turbine End of Faulkner, Strong and Kirk Turbocharger

Parameter SI Units EG Units


Lubricant Properties
Density ρl 850 kg/m3 0.031 lbm/in3
Specific heat capacity cl 2000 J/kg/oC 0.478 Btu/lbm/oF
o o
Supply temperature T0 87.8 C 190 F
Supply viscosity µ0 0.0066 Pa s 0.952 µreyn
β
o -1 o -1
Thermovisc. coeff. 0.031 C 0.017 F

Bearing Properties
Length L 0.057 m 2.24 in
Radial clearance Cb 7.11e-5 m 2.8 mils
Heat transfer coefficient H default W/m2/oC default hp/in2/oF
Bearing load Wb 916 N 206 lbf
Journal radius Rj 0.041 m 1.63 in
α
o -1 o -1
Journal C.T.E 1.10E-05 C 6.11E-06 F

Rotor Properties
Rotor weight W 1877 N 422 lbf
Overhung mass md 61.7 kg 136 lbm
Overhang Distance Ld 0.20 m 7.71 in
Max. Cont. Op. Speed ωMCOS 1047 rad/s 10000 RPM
Initial Mech. Unbalance Um 1.73e-4 kg m 0.24 oz in

If the heat transfer coefficient is not known (as is the case with the turbocharger bearings)

a default value can be calculated by using the following formula:

− 0.4
eq2. 40 H = (05
. )(255
  0.7
. ) ωR j  (µ 0 )− 0.2  2πR j
Equation eq2. 40 is based on an expression that was derived for tilting pad journal

bearings (Ettles, 1992).

During operation, it was observed that the turbocharger became unstable near

around 9900 RPM. Faulkner, Strong and Kirk initially thought that the instability was due

to the turbine wheel becoming loose at the high operation speeds. However, a careful

inspection of the turbine wheel position, before and after operation, indicated that the

44
wheel did not move while the turbocharger was running. Furthermore, a damped critical

speed analysis failed to justify the existence of a lateral critical speed near 9900 RPM. It

was finally concluded that the source of the instability was the thermal bowing of the rotor

shaft near the turbine end of the turbocharger.

After using the current thermal instability model to analyze the turbocharger, it

was found that the thermal instability was predicted to occur around 1009 rad/s (9640

RPM) near the turbine end of the turbocharger (Figure 2. 14).

Figure 2. 14 - Unbalance Curves from Current Analysis of Turbine End of the

Turbocharger studied by Faulkner et al.

45
Unlike the Keogh and Morton case, the turbocharger does not encounter criticals with low

damping in the given speed range . As a result, there are no large amplitude orbits to

produce a sudden instability hump. The monotonic increase in U can be explained by the

increase in speed which leads to increased viscous dissipation, higher ∆T values and more

incentive for thermal bending. Figure 2. 15 and Figure 2. 16 illustrate this trend of higher

speed leading to higher ∆T values.

Figure 2. 15 - Synchronous Orbit from Turbocharger (Turbine End) at 9000 RPM

46
Figure 2. 16 - Synchronous Orbit from Turbocharger (Turbine End) at 11000 RPM

The figures above also show that the orbits become more centered at higher speeds, i.e.

the center of the elliptical orbit approaches the bearing center (x/Cb = 0, y/Cb = 0) as the

rotor speed increases. A more centered orbit means that the hot spot and the cold spot

experience almost constant film thickness values, and such a configuration causes a

relatively steady thermal gradient to develop. In fact, this gradient would be the most

steady when the orbit is a circle that is completely centered. The Keogh and Morton case

show that large amplitude orbits tend to produce high thermal gradients which favor the

Morton Effect. If these high gradients are combined with a steady orientation, the worst

case scenario for the Morton Effect is achieved; large-amplitude, circular and centered

47
orbits are the most likely to be associated with the Morton Effect. This scenario can be

illustrated by considering Figure 2. 15 and some related orbits.

Figure 2. 15 shows a 9000 RPM synchronous elliptical orbit at the turbine end of

the turbocharger. The resultant (U) and threshold (Uthr) unbalances for this speed and

configuration are 0.36 oz in and 0.44 oz in respectively. If this orbit is forced to center

(Figure 2. 17) then the resulting ∆T is drastically increased from 0.87 oF to 6.07 oF. The

resultant unbalance is now 1.09 oz in. which is greater than the threshold value of 0.44 oz

in. As a result, the turbocharger is now experiencing the Morton Effect at 9000 RPM.

Figure 2. 17 - Forced 9000 RPM Turbine End Orbit: The Ellipse is Centered

By setting Ax = Ay = semi-major axis of the ellipse in Figure 2. 17, and by equating φy to

φx, the elliptical orbit can be changed into an equivalent circular one (Figure 2. 18). The

48
radius of the circular orbit is equal to the semi-major axis of the ellipse and the ∆T value is

increased. Now, the value of U is 1.16 oz in.

Figure 2. 18 - Forced 9000 RPM Turbine End Orbit: The Centered Ellipse is

Changed to a Circle with Radius = Semi-Major Axis of Ellipse

After increasing the diameter of the circular orbit by a factor of 10, Figure 2. 19 is

obtained. The ∆T value has risen to 10.85 oF and U is equal to 1.74 oz in. This entire

process shows that the Morton Effect becomes progressively worse as an orbit is

centered, then made to be more circular and finally enlarged.

49
Figure 2. 19 - Forced 9000 RPM Turbine End Orbit: The Centered Circular Orbit is

Enlarged by a Factor of Ten

50
2.10. Solution Sensitivity to the Threshold
Unbalance Force
The threshold unbalance force in the two cases studied above was 15% of the

rotor weight. Table 2. 3 shows the effect on the calculated instability speeds if different

threshold unbalance force magnitudes were chosen.

Table 2. 3 - Predicted and Actual Instability Speed Values (RPM) for Different

Threshold Unbalance Force Levels

Threshold Faulkner, Strong and Kirk


Unbalance Keogh and Morton Rotor Turbocharger (Turbine End)
Force
(% of Rotor Predicted Actual Predicted Actual
Weight)
5 >6853 9769-10371 >6420 >9900

10 >9603 9769-10371 >8320 >9900

15 10001-11521, 9769-10371 >9640 >9900

>12277

20 10383-11040 9769-10371 >10700 >9900

25 stable 9769-10371 >11510 >9900

It appears that a change of only 5% in the threshold unbalance force can change the

calculated instability speed by several hundred RPM. The solution is therefore sensitive to

the choice threshold unbalance force. From the data shown in Table 2. 3, the optimum

threshold unbalance force seems to be 15% of the rotor weight. This choice of unbalance

force is the one which gives data that most corresponds with the Keogh and Morton

instability range. In addition, this force magnitude predicts an instability speed that lies

51
close to the observed thermal instability speed at the turbine end of the Faulkner, Strong

and Kirk turbocharger. As a result, the 15% unbalance force level was utilized in the

threshold instability criterion.

52
3. The Morton Effect in Rotors with Tilting
Pad Journal Bearings

The procedure for developing a Morton Effect model for rotors with tilting pad

journal bearings (TPJB) is similar to that for plain journal bearings. In fact, the steps for

calculating the initial mechanical unbalance (Um), the thermal unbalance (Ut) and the

instability threshold unbalance (Uthr), are identical. In addition, the computer flowchart,

which is given in chapter 2, is also applicable to rotors supported by tilting pad journal

bearings. It would therefore be redundant to repeat these sections in this chapter.

Tilting pad journal bearings are generally more stable than plain journal bearings

because the tilting pads theoretically set the bearing cross-coupled stiffness coefficients to

zero (Kirk, 1998). This property minimizes the tangential forces acting on the shaft and

thus stabilizes the system. The fixed geometry of the plain journal bearings does not allow

the cross-coupled stiffness coefficients to be set to zero and, as a result, plain journal

bearings tend to be less stable than their TPJB counterparts.

However, the presence of tilting pads, which are free to move independently of

each other, increases the complexity of the modeling. A numerical solution will therefore

have to be adopted. Furthermore, the film thickness expressions for TPJBs have not been

explicitly derived in any current work of literature. An attempt will be made to develop

53
film thickness expressions for the cylindrical pivot TPJBs and extend the modeling to

include spherical pivot TPJBs.

3.1. Film Thickness for Tilting Pad Journal


Bearings
Tilting pad journal bearings can either be of a load-on-pad (LOP) or of a load-

between-pads (LBP) configuration. These two configurations are schematically shown in

Figure 3. 1 and can adopt a variety of pivot types.

Lubricant

Journal

Pad

Pivot

LOP LBP

Figure 3. 1 - Schematic Diagrams of Tilting Pad Journal Bearings Showing Load-

on-Pad and Load-between-Pads Configurations

Regardless of the configuration, a single pad of a TPJB with a cylindrical pivot can be

illustrated as shown in Figure 3. 2. Most current TPJBs carry 4 or 5 pads (Figure 3. 1),

but only one pad is shown in Figure 3. 2 in order to avoid congesting the drawing.

54
Y

bearing
M
tp
δ
N
h
J pad
Rb
Rp
lubricant θc
θ
Ob θp
X
e
Op Oj

journal
Op’

Figure 3. 2- Schematic Diagram of a Cylindrical Pivot TPJB

55
This pad is depicted before (dotted line) and after (solid line) tilting it through an angle δ.

The center of the bearing is represented by Ob and the center of the journal is Oj . Points

Op and Op’ respectively refer to the pre-tilt and post-tilt positions of the pad center of

curvature. Rb corresponds to the bearing radius while Rj denotes the journal radius. The

distance between the pivot point, M, and the surface of the pad is tp, and the pad radius of

curvature is Rp. This pad is located at an angular position θp relative to a stationary

coordinate system with origin Ob. θc represents the angle to the line-of-centers which

connects Ob to Oj. The distance between Ob and Oj is the journal eccentricity, e. The film

thickness, h, is measured at an angle θ with respect to the given coordinate system. This

circumferential angle, θ, is different from the angle ξ defined for the plain journal bearing.

In the plain journal bearing case, ξ was defined relative to the moving line-of-centers, with

the clockwise direction being positive. On the other hand, the TPJB case defines θ with

respect to Ob and a stationary coordinate system, with the counter-clockwise direction

being positive. This new definition enables a more conventional coordinate system to be

used which simplifies the numerical analysis. However, it can be shown that if the line-of-

centers position is taken into account, and the pad effects are neglected, the TPJB film

thickness expression can be reduced to give an equivalent plain journal bearing one.

Some of the triangles in Figure 3. 2 can be extracted and represented as follows:

56
M

δ
Y
Rb+tp
Rp+tp

(θp+δ) N
θp
Ob γ X Y ObJ+h λ
J r ξ
Op’
ο Rp

π−(θc−θ) θ
Rj Ob γ X
r κ
Ob Op’
e ι
Oj

(a) (b) (c)

Figure 3. 3 - Triangles from Diagram of The Cylindrical Pivot TPJB

From Figure 3. 3a:

Rj e O J
eq3. 1 = = b
[
sin π − (θc − θ) ] sin ο sin ι

Solving for ο from eq3. 1 gives:

eq3. 2 ο = sin − 1 

(
e sin θ c − θ  )
Rj 
  

Solving for ι from eq3. 1 gives:

57
   
    − 1  e sin θ c − θ 
eq3. 3 ι = π − π −  θ c − θ + ο = θ c − θ − sin  
   Rj 
 

Hence,

O J=
R j sin ι
=

Rj
 −
sinθ c − θ − sin 
1
 e sin θ − θ
c ( ) 
eq3. 4 
b
( )
sin θ c − θ sin θ c − θ ( 

)  R j 

Since e << Rj, the sin-1 term approximates to its argument. Therefore, eq3. 4 can be

reduced to the following expression:

Rj   e sin θ − θ
( ) 
eq3. 5 ObJ =

(
sin  θ c − θ −  ) c
(
sin θ c − θ 

)  R j

 

Expanding this expression using trigonometric identities:

Rj   e sin θ − θ
( )  − cos θ − θ sin e sin(θ c − θ) 
eq3. 6 O b J =

( )
sin θ c − θ cos
c
( c )  R 
( )
sin θ c − θ 
  Rj 

 j  

Applying the small angle approximations cos(x) ~ 1 and sin(x) ~ x to eq3. 6 gives the

following upon simplification:

eq3. 7 O b J = R j − e cos θ c − θ ( )
The cosine rule can be used on Figure 3. 3b to give:

1
 2 2  2
eq3. 8 r =  R p + t p  +  R b + t p  − 2 R p + t p   R b + t p  cosδ
 
 

58
Typical values for the parameters in eq3. 8 are tp = 0.5 in. (1.27 cm), Rp = 2.006 in. (5.095

cm), Rb = 2.003 in. (5.088 cm) and δ = 0.05 degrees. These values give r = 0.0037 in.

(94.3 µm) or r/Rp = 0.0018 which is much less than 1. From Figure 3. 3c:

Rp r O N
eq3. 9 = = b
sin(θ + γ ) sin λ sin κ

By rearranging eq3. 9 and applying the approximation r/Rp << 1 yields:

(
 r sin θ + γ 
)
 ≈ r sin θ + γ( )
eq3. 10 λ = sin −1 

Rp 
Rp

Also,

r sin(θ + γ )
eq3. 11 κ = π−θ− γ −λ = π−θ− γ −
Rp

Equation eq3. 9 can now be further rearranged to produce:


ObN =
R p sin κ
=

Rp
sinπ − θ − γ −
r sin θ + γ ( ) 
eq3. 12 
( )
sin θ + γ sin θ + γ 

( ) Rp 

Since sin(π-x) = sin(x), this equation can be re-written and expanded to give:


Rp 

 r sin θ + γ ( )  
 r sin θ + γ( ) 
eq3. 13 ObN = (
sin θ + γ cos ) + cos (
θ + γ )
sin 
( )
sin θ + γ 
  Rp 


 Rp  

Applying the r << Rp and small angle approximations gives :

eq3. 14 O b N = R p + r cos(θ + γ )

59
Figure 3. 3b indicates that γ = π − θ p − δ − ξ . If this expression for γ is substituted into

equation eq3. 14, then the following equation ensues:

eq3. 15 O b N = R p − r cos(δ + ξ) cos θ − θ p  − r sin(δ + ξ) sin θ − θ p 

From Figure 3. 3b,

r Rb +tp
eq3. 16 = ⇒ r sin ξ =  R b + t p  sin δ
sin δ sin ξ

Therefore,

1
 1/ 2  2 
eq3. 17 r cos ξ = r 1− sin 2 ξ
 
( )
= r − Rb + tp 2 sin 2 δ 2
 

Substituting eq3. 8 into eq3. 17 and simplifying gives:

eq3. 18 r cos ξ =  R p + t p  −  R b + t p  cos δ

Using eq3. 16 and eq3. 18:

eq3. 19 r cos(δ + ξ) = r cos δ cos ξ − r sin δ sin ξ =  R p + t p  cos δ −  R b + t p 

eq3. 20 r sin(δ + ξ) = r sin δ cos ξ + r cos δ sin ξ =  R p + t p  sin δ

Since δ is very small (~ 0.05 deg or 8.73e-4 rad), sin(δ) ~ δ, and cos(δ) ~ 1. Therefore,

equations eq3. 19 and eq3. 20 can be re-written as follows:

eq3. 21 r cos(δ + ξ) = R p − R b

60
eq3. 22 r sin(δ + ξ) = δ R p + t p 

Substituting eq3. 21 eq3. 22 and eq3. 22 into eq3. 15 gives:

eq3. 23 O b N = R p −  R p − R b  cos θ − θ p  − δ R p + t p  sin θ − θ p 

From Figure 3. 3c, it can be deduced that the film thickness is:

eq3. 24 h =O b N −ObJ

Combining eq3. 7, eq3. 23 and eq3. 24 gives an expression for the film thickness of a

cylindrical pivot TPJB:

eq3. 25 ( )
h =  R p − R j  + e cos θ c − θ −  R p − R b  cos θ − θ p  − δ R p + t p  sin θ − θ p 

Other researchers have published expressions for the film thickness of a cylindrical pivot

TPJB and some of these formulae are listed below (the symbols have been changed to

correspond with the nomenclature used in this dissertation).

Monmousseau et al., 1997 (the pad deformation effects have been neglected):

eq3. 26 h = C p − C p m cos θ − θ p  −  R j + t p  δ sin θ − θ p  + e cos θ c cos θ + e sin θ c sin θ

Ha et al., 1995:

eq3. 27 h = C p + e cos(θ c − θ) − C p m cos θ − θ p  − δR j sin θ − θ p 

Kim et al., 1994:

eq3. 28 h = C p + e cos θ c cos θ + e sin θ c sin θ −  C p − C b  cos θ − θ p  − δR j sin θ − θ p 

61
where the radial pad clearance, Cp = Rp-Rj, the radial bearing clearance, Cb = Rb-Rj and the

preload factor, m = 1 - Cb/Cp. The above equations can be compared by using some

typical values which are given in Table 3. 1.

Table 3. 1- Input Data for Comparison of Film Thickness Expressions

PARAMETER VALUE

Rb Radius of bearing 2.003 in. (5.088 cm)

Rj Radius of journal 2.0 in. (5.080 cm)

Rp Inner radius of curvature of pad 2.006 in. (5.095 cm)

tp Pad thickness 0.5 in. (1.27 cm)

∆p Angular dimension of pad 56 deg

θp Pivot angles of pads 54, 126, 198, 270, 342 deg

θc Angle for line-of-centers (Ob,Oj) 90 deg

δ Pad tilt angle 0.05 deg

Npads Number of pads 5

fp Fractional angular position of pivot 0.5

ε Eccentricity ratio (e/(Rb-Rj) 0.6

The fractional angular position of the pad pivot, fp = (distance between the leading edge of

the pad and the pivot )/(total angular length of the pad, ∆p). The data in Table 3. 1 was

used to generate Figure 3. 4. This figure shows that h from the current research compares

fairly well with the film thickness equations from the other studies. In fact, along with

Monmouseau et al., this research gives the most conservative estimate of the film

thickness, i.e. the lowest values of h. This is because the current research and the work

62
done by Monmousseau take into account the pad thickness, tp. This additional

consideration causes the film thickness to be lower, especially at pad 4 which is the most

heavily loaded pad. As a result, the current research allows for a safer design which

would be less likely to suffer from the journal destroying the pad babbitt due to an

overestimation of the film thickness value.

Current Research
Monmousseau et al.
6 Ha et al.
Kim et al.

5
Film Thickness, h (mils)

Pad 1 Pad 2 Pad 3 Pad 4 Pad 5

0
0 50 100 150 200 250 300 350 400
theta (degrees)

Figure 3. 4 - Comparison of Film Thickness Expressions for a 5-Pad Cylindrical

Pivot TPJB

It should be noted that, if Rp = Rb and δ = 0, equation eq3. 25 will reduce to give

an equivalent expression for the film thickness of a plain journal bearing. In other words,

63
eq3. 25 consists of the plain journal bearing film thickness ( (Rb - Rj) + ecos(θc-θ) = Cb[1+

εcos(θc-θ)]) plus a contribution from the pre-loaded pad ( ( Rp - Rb)[1-cos(θ-θp)] ) plus a

contribution due to the tilting of the pad ( -δ(Rp+tp)sin(θ-θp) ).

Some TPJBs contain spherical pivots which have more degrees of freedom than

cylindrical pivots. However, any pad-tilting about the x-axis or y-axis usually leads to

journal misalignment which could create an unstable unbalance response (Vance, 1988).

Therefore, it will be assumed that the spherical pivots only tilt about the z-axis and will

thus have the same film thickness expression as the cylindrical pivot TPJBs.

3.2. Static Equilibrium Journal Position


The additional degrees of freedom from the pads and the complexity of the film

thickness expression make it necessary for a numerical solution to be used to find the

journal position. At the inception of this project, it was thought that the complicated

geometry of the TPJB domain would be difficult to model accurately using the rectangular

grid of the finite difference method. As a result, the finite element method was chosen to

deal with this complex geometry. In retrospect, however, finite difference techniques

would have probably sufficed.

In order to determine the static equilibrium journal position, the pressure

distribution must first be obtained by numerically solving the Reynolds’ equation. This

pressure must then be used to balance the forces on the journal and the moments on each

pad in the TPJB so that the static equilibrium position of the journal can be determined.

64
3.2.1. Reynolds’ Equation
A local coordinate system can be set up in the region between the pad and the

journal surfaces as shown in Figure 3. 5.

Pad Surface

h
Lubricant
y

uj = ωRj Journal Surface


z

Figure 3. 5 - Coordinate System for Reynolds’ Equation

In 1886, Osborne Reynolds derived the following equation from the Navier-Stokes

equations and a coordinate system similar to the one shown in Figure 3. 5.

   
∂  h 3 ∂P  ∂  h 3 ∂P  ∂h
eq3. 29  +   = 6u j
∂x  µ ∂x  ∂z  µ ∂z  ∂x

where P = pressure and µ = lubricant viscosity. The derivation of equation eq3. 29 has

also been done by Cameron (Cameron, 1966) and will not be repeated here. Instead, the

65
assumptions made in the derivation of equation eq3. 29, and their implications to this

research, will be discussed.

The following assumptions (Cameron, 1966) were made in order to derive the

Reynold’s equation:

(1) Body forces are neglected. This means that gravitational, magnetic and electrical

forces are assumed not to be present. Such an assumption is true for most of the

mineral oils which are used as lubricants in bearings.

(2) Pressure is constant through the film thickness. As a result, the pressure gradient in

the y-direction is zero and no lubricant motion is assumed to occur in this direction. If

the lubricant is highly compressible, e.g. a gas, then this assumption will not be true.

(3) Bearing surface curvatures >> h. Displacement in the x-direction can therefore be

approximated as x ~ Rj θ, where θ is defined in Figure 3. 2. This approximation also

assumes that Rb ~ Rj, which is a reasonable statement as indicated by the values in

Table 3. 1.

(4) No slip at the boundaries. Hence, u(y=0) = uj and u(y=h) = 0.

(5) Newtonian lubricant. A Newtonian lubricant has a shear stress (τ) which is equal to

the shear strain rate (e.g. du/dy) multiplied by the lubricant viscosity (µ).

(6) Laminar flow. The Reynolds number (p. 441, Cameron, 1966) for bearings can be

defined as Re = ωRjCbρl/µ and the critical Reynolds’ number which demarcates the

transition to turbulence is given by Rec = 41.3 (Rj/Cb)1/2. If typical values of ω =

10,000 RPM (1047 rad/s), Rj = 2 in. (5.08 cm), Cb = 3 mils (76.2 µm), ρl = 0.031

lbm/in3 (860 kg/m3), µ = 2.0 µreyn (13.8 cP or 13.8 mPa s) are used then Re = 253 and

66
Rec = 1066. Therefore, most continuous fluid film bearings typically operate under

laminar flow conditions.

(7) Neglect fluid inertia. This assumption allows the mass x acceleration term in the

Navier-Stokes equations to be neglected. Consequently, the pressure and shear stress

gradients can be equated to provide a starting point for the development of the

Reynolds’ equation.

(8) Constant viscosity across the film thickness. This simplifies cross-film integration

when calculating velocity profiles.

3.2.2. Pressure Distribution


Reynolds’ equation can be solved over the domain of each pad surface in the TPJB

to give the pressure distribution. A typical meshed pad surface with essential pressure

boundary conditions is shown in Figure 3. 6.

z (axial)
Mesh of
Quadrilateral
θ,x (circumferential) Elements

z = +L/2, P = P0
θ=
θ=

θ e,
θ s,

P=
P=

P0
P0

z = -L/2, P = P0

Figure 3. 6 - Pad Surface Showing Pressure Boundary Conditions

67
where θs = start angle of the pad, θe = end angle of the pad and L = axial pad length. The

above domain can be meshed using quadrilateral elements which will be discussed in detail

shortly. It should be noted that the h-version (instead of the p-version) of the finite

element method will be chosen for simplicity. In other words, lower order basis or

interpolation functions will be used and convergence will be achieved by refining the mesh.

The p-version uses a given number of elements and achieves convergence by increasing

the order of the polynomials used as basis functions. Generally, the p-version converges

faster, but is slightly more complicated to implement than the h-version.

In the finite element method, basis functions are used to approximate the exact

solution to a differential equation. These basis functions are defined over simple elements

which comprise a mesh of the particular domain. Accompanying these basis functions are

unknown coefficients which can be obtained by setting the weighted integral of a residual

to be equal to zero (Reddy, 1993). The residual is obtained from the original governing

equation. Such a weighted integral statement is embodied in the weak formulation of

Reynolds’ equation, where the weight or test function is denoted by w:

  3   3  
 ∂  h ∂P  ∂  h ∂P  ∂h 
eq3. 30 ∫∫ w    +   − 6u j dA = 0

element  x  µ ∂ x  ∂ z  µ ∂z  ∂ x 

Using the following identity (Reddy, 1993),

∂f ∂w ∂( wf ) ∂w
eq3. 31 ∫∫ w dA = − ∫∫ f dA + ∫∫ dA =− ∫∫ f dA + ∫ wfn x dl
∂x ∂x ∂x ∂x
equation eq3. 30 can be written as:

68
 3 3 
 h ∂P ∂w + h ∂P ∂w dA =
∫∫ 
µ ∂x ∂x µ ∂z ∂z 
element  
eq3.32
 3 3 
 h ∂P n + h ∂P n dl −  ∂h 
∫ w  µ ∂x x µ ∂z z  ∫∫ w 6u j dA
 ∂x 
boundary   element 
At this point, the basis function approximations (eq3. 33) and the Galerkin approximation

(eq3. 34 - weight function = basis function) can be introduced:

NPE
eq3. 33 P ≈ ∑ P jφ j
j= 1

eq3. 34 w =φ i

where NPE = number of nodes per element and φk is used to represent the basis function.

The weight function, w, can also be thought of as a variation of the primary variable, P. If

P is specified on the boundary the variation of P, and consequently w, will be zero. From

Figure 3. 6 it can be seen that P is specified as the lubricant supply pressure (P0) on all

four boundaries of the pad surface. Hence, the boundary integral on the right hand side of

eq3.32 evaluates to zero. Noting this observation and substituting eq3. 33 and eq3. 34

into eq3.32 gives:

NPE  h 3  ∂φ i ∂φ j ∂φ i ∂φ j  

 ∂h 
eq3. 35 ∑  ∫∫ + dA P j =− ∫∫ φ i 6u J dA
 
j= 1 element µ  ∂x ∂x ∂z ∂z   element  ∂x 

Equation eq3. 35 can be re-written as:

eq3. 36

NPE
∑ KE ijP j = FE i
j =1
h 3  ∂φ i ∂φ j ∂φ i ∂φ j   ∂h 
where KE = ∫∫ + dA and FE = − ∫∫ φ i 6u J dA
ij µ  ∂x ∂x ∂z ∂z  i  ∂x 
element element 

69
KE refers to the element “stiffness” matrix and FE is the element load or “force ” vector.

Equation eq3. 36 constitutes the foundation for the finite element method.

Before the integrals contained in KEij and FEi can be evaluated, the basis functions

have to be defined. These linear basis functions will be used to facilitate numerical

integration over elements which are similar to the one shown in Figure 3. 7.

4 3

dz η

1 2

dx = Rj dθ

Figure 3. 7 - A 4-Node Isoparametric Quadrilateral Element

The linear basis or interpolation functions and their derivatives can be expressed as

follows:

1
(a ) φ = 1+ ξ cjξ 1+ ηcj η
j 4
∂φ j ξ cj

eq3. 37 (b) = 1+ ηcjη
∂ξ 4  

∂φ j ηcj
(c) = 1+ ξ ξ
∂η 4  cj 

The values for some of the terms in equation eq3. 37 are given in Table 3. 2:

70
Table 3. 2 - Basis Function Parameters

Local Node # ξcj ηcj m ξ =ξm η =ηm wm

1 -1 -1 1 -1/√3 -1/√3 1

2 +1 -1 2 +1/√3 -1/√3 1

3 +1 +1 3 +1/√3 +1/√3 1

4 -1 +1 4 -1/√3 +1/√3 1

The chain rule can now be used to express the isoparametric derivatives (from eq3. 37) in

terms of a Jacobian matrix and Cartesian derivatives:

 ∂φ   ∂x ∂z  ∂φ 
   
 ∂ξ   ∂ξ ∂ξ  ∂x 
 = 
∂z  ∂φ 
eq3. 38
 ∂φ   ∂x 
 ∂η   ∂η ∂η  ∂z 
  

The Jacobian matrix (JA) can be approximated in terms of the derivatives of the basis

functions:

 NPE ∂φ j NPE ∂φ j 
 ∂x ∂z   ∑ x ∑ zj 
   j ∂ξ ∂ξ 
∂ξ ∂ξ   j=1 j= 1
eq3. 39 JA =  = 
∂x ∂z   NPE ∂φ j NPE ∂φ j 
   
 ∂η ∂η   ∑ x j ∂η ∑ zj
∂η 
 j=1 j= 1 

Hence, the Cartesian derivatives which are required by eq3. 36 can be obtained from eq3.

38 and eq3. 39:

71
∂φ j −1
NPE NPE ∂φ j 
 ∂φ   ∑ x ∑ zj   ∂φ 
   j ∂ξ ∂ξ   
 ∂x   j= 1 j =1   ∂ξ 
eq3. 40  =   
 ∂φ  NPE ∂φ j NPE ∂φ j   ∂φ 

 ∂z   ∑ x j ∑ zj  ∂η 
 j= 1 ∂η j =1 ∂η   

The integration in eq3. 36 is usually done numerically by Gaussian Quadrature which can

be illustrated as follows:

1 1 NG
eq3. 41 ( )
∫∫ f ( x, z)dA = ∫∫ f ( x, z)dxdz = ∫ ∫ f ξ, η JA dξdη = ∑ w m f  ξ m , η m  JA
element element −1 −1 m =1

where NG = number of Gauss points = 4, and the values for the other parameters are

given in Table 3. 2. In order to avoid the use of two loops in the computer program, the

traditional double summation, which is implied by the integration, has been collapsed into

a single equivalent summation in the last term of eq3. 41.

The integrals in eq3. 36 can now be evaluated using eq3. 41 and the resulting

element matrices can be assembled to give global matrices. The boundary conditions can

now be applied, and these global matrices can then be solved for nodal pressure values

using a banded matrix solver. All of these operations are executed by the program listed

in Appendix B.

3.2.3. Forces and Moments


The pressure distribution can be integrated over an infinitesimal pad area to give a

force which can be translated into a moment. For static equilibrium to be achieved, these

moments must be equal to zero:

72
L/ 2 xe     
eq3. 42 MOMENTi = ∫ ∫  R j + t p   θ − θ p  P cos θ − θ p  dxdz = 0, i =1,2,..., N pads
− L/ 2 xs

Equation eq3. 42 is an approximation which was derived from Figure 3. 2 and it can be

numerically evaluated using eq3. 41. The symbols xs = Rjθs and xe = Rjθe. Like the pads,

the journal is also free to move. Static equilibrium requires that the X and Y forces on the

journal balance to zero:

N pads  L / 2 x e 
 
∑ FX = ∑  ∫ ∫ ( P cos θ)dxdz = 0
i = 1 − L / 2 x s 
i
eq3. 43
N pads  L / 2 x e 
 
∑ FY = ∑  ∫ ∫ ( P sin θ) dxdz − Wb = 0
i =1 − L / 2 x s 
i

where Wb = bearing load. Equations eq3. 42 and eq3. 43 can be solved simultaneously by

using the Newton-Raphson method for systems. This method solves the system of

equations shown below:

 MOMENT
 1  0
 
 M  M 
   
⇒ F( X) = 0
MOMENT  ~ ~ ~
eq3. 44  N pads  = 0
   
 ∑ FX  0
   
 ∑ FY  0
~
where X consists of the tilt angles (δ) for each pad, the journal eccentricity (e) and the

angle to the line-of-centers (θc). At the foundation of the Newton-Raphson method lies

the following Taylor series expansion:

~
~  ~  ~ ~  ∂ F
  X
~ ~  ~
eq3. 45 F X k + 1  ≈ F X k  + ~  k +1 − X k  = 0
∂X ~
Xk

73
~
~ ∂F
( )
~ ~ ~ ~
Defining J = ~ and Y = X k +1 − X k yields after rearranging eq3. 45:
∂X X
~
k

~ −1 ~ ~
eq3. 46
~ ~
( )
Y =−  J  F X k
 
~
Equation eq3. 46 can be solved for each successive kth iteration until Y becomes less than

some pre-defined tolerance. At this point the iterations have converged to give a solution
~
to eq3. 45 which, in turn, solves eq3. 42 and eq3. 43. The final result yields a X vector

which contains the equilibrium tilt angles for the pads and the e and θc which define the

static equilibrium journal position.

The FEM program was checked for convergence and validated by comparing the

results with those from other studies. The first comparison was done with data presented

by Someya (p. 224-225, Someya, 1989) which is shown in Table 3. 3.

74
Table 3. 3 - Input Data for Bearing Number 10 (5-Pad TPJB), Test Number 10

(Someya, 1989)

PARAMETER VALUE

L Axial pad length 1.97 in. (5.0 cm)

tp Pad thickness 0.5 in. (1.27 cm)

Cp Radial pad clearance 8.50 mils (216 µm)

∆p Angular dimension of pad 60 deg

θp Pivot angles of pads 54, 126, 198, 270, 342 deg

m Preload factor 0.5

fp Fractional angular position of pivot 0.5

Rj Radius of journal 1.97 in. (5.0 cm)

ω Angular journal speed 10020 RPM (1049 rad/s)

Wb Bearing load 2203 lbf (9800 N)

P0 Lubricant supply pressure 19.1 psi (0.132 MPa)

µ Lubricant viscosity 2.32 µreyn (16 mPa s)

ρl Lubricant density 0.03 lbm/in3 (865 kg/m3)

Someya obtained an experimental static eccentricity ratio of 0.63 (p. 224 Someya, 1989),

but the attitude angle (ψa0 = θc-π/2) was not given. He also developed a numerical model

which yielded ε0 = 0.58 and ψa0 = 0.0 deg. The following data is from the program which

75
was developed in the last few sections. NEX = number of elements in the x-direction and

NEZ = number of elements in the z-direction.

Table 3. 4 -Computed ε0 and ψ a0 as a Function of Mesh Size for Someya’s Bearing

NEX NEZ (axial)

(circ.) 5 10 15

5 0.66 0.65 0.65

0.021 deg 0.021 deg 0.021 deg

10 0.64 0.64 0.64

0.05 deg 0.05 deg 0.05 deg

15 0.64 0.64 0.64

0.05 deg 0.05 deg 0.05 deg

From Table 3. 4, it appears that the numerical solution has converged to values of ε0 =

0.64 and ψa0 = 0.05 deg for NEX = 10 and NEZ = 5 (in order to maintain symmetry, 6

elements will be used in the z-direction along with the 10 elements in the x-direction).

Furthermore, the solution of ε0 = 0.64 is closer to the experimental value than the solution

obtained by Someya. It should be noted that most TPJBs, which are symmetric about the

Y-axis, have a static equilibrium position that lies on the negative Y-axis. This means that

the attitude angle should be close to zero degrees, as is evident from the data in Table 3.

4. The slight deviation from the theoretical zero degrees in Table 3. 4 could be due to a

numerical truncation error.

76
Another TPJB was investigated by Josiah Knight (Knight, 1990). This 5-pad

TPJB was loaded in a load-between-pad or LBW configuration, which means that there is

no pivot at θp = 270 deg as in the Someya case (load-on-pad or LOP configuration). Data

for Knight’s bearing is given in Table 3. 5.

Table 3. 5 - Input Data for a 5-Pad LBP TPJB (Knight, 1990)

PARAMETER VALUE

L Axial pad length 2.24 in. (5.7 cm)

tp Pad thickness 0.5 in. (1.27 cm)

Cp Radial pad clearance 4.72 mils (120 µm)

∆p Angular dimension of pad 54 deg

θp Pivot angles of pads 18, 90, 162, 234, 306 deg

m Preload factor 0.5

fp Fractional angular position of pivot 0.5

Rj Radius of journal 1.5 in. (3.8 cm)

ω Angular journal speed 9000 RPM (942 rad/s)

Wb Bearing load 1000 lbf (4450 N)

µ Lubricant viscosity 1.02 µreyn (7 mPa s)

ρl Lubricant density 0.03 lbm/in3 (850 kg/m3)

The results from the current program yielded ε0 = 0.62 and ψa0 = 0.005 deg, while

Knight’s program predicted ε0 = 0.60 and ψa0 = 0.0 deg. The actual experimental values

were ε0 = 0.64 and ψa0 = -2.2 deg. The negative non-zero value of ψa0 in the

77
experimental data indicates that there was some sort of asymmetry introduced into the

bearing, e.g. friction between the pad and the pivot could cause some of the pads to stick

in a fixed orientation which would force the journal to conform to that pre-defined

position.

3.3. Synchronous Orbit


The process of defining the synchronous orbit for the TPJB case is the same as for

the plain journal bearing rotors. In fact, there is no ξ-coordinate system in the TPJB case

and this makes it easier to obtain the locations of the hot and the cold spots. These spots

can be located by using the equations that are given in chapter 2 for xH, yH, xC and yC.

Unfortunately, the stiffness and damping coefficients (which are required by VT-

FAST) for tilting pad journal bearings do not have a convenient closed form solution as do

those for the plain journal bearings. As a result, a numerical computation is required and

this calculation involves the following steps:

(1) Reynolds’ equation is first solved and the resulting data is used to establish an

equilibrium position for the journal and pads.

(2) A perturbation is then applied to the journal. This perturbation can be a displacement

or a velocity increment in either the X or the Y direction.

(3) The pads are then allowed to adjust to the new journal position by obtaining tilt angles

that set the pad moments to zero.

(4) The differences between the equilibrium forces and the new X and Y forces on the

journal are then calculated. Dividing these force differences by the perturbation

increment leads to the stiffness and damping coefficients.

78
Before a perturbation can be applied to the journal, the Reynolds’ equation must

be modified in order to accommodate this perturbation. Hence, equation eq3. 29 has to be

rewritten as follows:

   
∂  h 3 ∂P  ∂  h 3 ∂P  ∂h ∂h
eq3. 47   +   = 6u j +12
∂x  µ ∂x  ∂z  µ ∂z  ∂x ∂t

The transient term in eq3. 47 is sensitive to the perturbations and can be evaluated by

taking the partial derivative of the film thickness with respect to time. Equation eq3. 25

(the film thickness expression) contains the term ecos(θc-θ) which can be expanded to give

ecos θc cos θ +esin θc sin θ = -Xj cos θ - Yj sin θ. Now, the only terms in eq3. 25 which

are time-dependent are Xj (the X-coordinate of the journal center Oj), Yj (the Y-

coordinate of the journal center Oj) and δ, the tilt angle of the pad. Therefore, the

transient part of the last term in eq3. 47 can now be expressed as:

∂h  & & sin θ + δ&  R + t  sin θ − θ  


eq3. 48 =−  X j cos θ + Y j  p p  p  
∂t 

As a result of this new term, the finite element formulation of the Reynolds’ equation has

to be reformulated as follows:

eq3. 49

NPE
∑ KE ij P j = FE i
j= 1
h 3  ∂φ i ∂φ j ∂φ i ∂φ j   ∂h ∂h 
where KE = ∫∫ + dA and FE = − ∫∫ φ i 6u J −12 dA
ij µ  ∂x ∂x ∂z ∂z  i  ∂x ∂t 
element element 

79
The tilt angle velocity term, δ& , is typically zero if the above-mentioned method for

obtaining the stiffness and damping coefficients is used. However, if a pad perturbation

technique is used (Allaire et al., 1981; Nicholas et al, 1977) then this term will be non-

zero. For the purposes of this research, δ& will be set to zero. The stiffness and damping

coefficients for the TPJB can now be defined:

∆Fx ∆Fx ∆Fy ∆Fy


K = K =≈0 K = ≈0 K =
xx ∆X xy ∆Y yx ∆X yy ∆Y
j j j j
eq3. 50
∆F ∆F ∆Fy ∆Fy
C xx = x C xy = x ≈ 0 C yx = ≈ 0 C yy =
&
∆X &
∆Y &
∆X &
∆Y
j j j j

The cross-coupled coefficients Kxy, Kyx, Cxy and Cyx are usually close to zero and

negligible compared with the other direct coefficients (Shapiro and Colsher, 1977). This

implies that the TPJBs are more stable than the plain journal bearings. Cross-coupled

coefficients tend to produce tangential follower forces (Chapter 1, Vance, 1988) which

augment rotor whirl and eventually drive the system unstable. As a result, bearings with

lower cross-coupling terms, such as TPJBs, are generally more stable than bearings with

higher cross-coupled coefficients, e.g. plain journal bearings. In the current research, it

will be assumed that the TPJB cross-coupled coefficients are effectively zero.

It should be noted that the stiffness and damping coefficients in equation eq3. 50

can be made dimension-less by re-writing them as follows:

Cp
K′ = K
Wb
eq3. 51
ωC p
C′ = C
Wb

80
where K represents a generic stiffness coefficient and C represents a generic damping

coefficient. By using Someya’s bearing (Table 3. 3) as the input to the Appendix B

program, the following figure can be obtained:

12

Someya-Exp.
Someya-Theory
10 Current Theory

8
Dimensionless Value

0
K'xx K'yy C'xx C'yy
Coefficient

Figure 3. 8 - Comparison between Calculated and Actual Dimensionless Stiffness

and Damping Coefficients from Someya’s 5 -Pad Bearing

Figure 3. 8 shows that the current model gives fair agreement with experimental and

theoretical data from Someya (p. 158,224, Someya, 1989).

81
3.4. Temperature Distribution
The energy equation from chapter 2 can also be adapted for use with the TPJBs.

However, the complexity of the TPJB film thickness expression and the use of Fortran

code makes it necessary to employ a numerical technique to solve this differential

equation. Euler’s method, which is a finite difference formulation, can be used to solve

this energy equation:

 
  2H    2 HT 2µωR 2j 
    amb
eq3. 52 Ti +1 = 1− ∆θ   i
T + ∆θ + 
  ρ l c l ωh  i   ρ c ωh
l l ρ c h 2 
  l l  i

where ∆θ = increment in pad angle. Equation eq3. 52 can now be solved by specifying an

inlet temperature at the leading edge of each tilted pad. The inlet temperature (Tin) at one

of the pads is taken to be the sum of the lubricant supply temperature and the estimated

average temperature rise (∆Tav) multiplied by a scaling ratio:

h
eq3. 53 Tin = T0 + min ∆Tav
h in

where hmin = minimum inlet film thickness and hin = inlet film thickness for current pad.

This relationship assumes that higher inlet film thicknesses are associated with lower inlet

temperatures. It should be noted that the mixing of the lubricant in the pockets between

the pads and convective heat transfer at the pad boundaries were neglected. The bearing

heat transfer coefficient will be determined from the following equation (Ettles, 1992):

− 0.4
eq3. 54 . u 0j .7 µ 0− 0.2  R j∆ p 
H = 255

82
where H = heat transfer coefficient in Wm-2K-1, uj = journal surface speed in m/s, µ0 =

lubricant supply viscosity in Pa s and Rj∆p = pad arc length in m.

A pertinent question is: Why wasn’t the finite element method used to solve a

more complex form of the energy equation for the temperature distribution? The answer

is an attempt was made to obtain an FEM model of the temperature field. However,

implementation of the Galerkin approximation lead to unstable oscillations which was in

concordance with the observations of Abdel-Hadi et al. (1985) who did work on

advection-dominated diffusion equations. The recommended remedy to this problem is

the SUPG (Streamline Upwinding Petrov-Galerkin) model (Abdel-Hadi et al., 1985 and

Kim et al., 1994). This model accounts for the fluid moving upstream, and uses weighting

functions which are chosen to be different from the basis functions as is done in the

classical Bubnov-Galerkin (Galerkin) model. Unfortunately, efforts to implement the

SUPG model were unsuccessful and the more simple model summarized in eq3. 52 was

selected. This simpler model gives a reasonable estimate of the temperature distribution

that suffices for the purposes of this project.

Equation eq3. 52 is solved by the program in Appendix B and a TPJB studied by

Knight (Knight, 1990) was chosen to check for convergence and accuracy. The input data

for this TPJB is given in Table 3. 6.

83
Table 3. 6 - Input Data for a 5-pad LBP TPJB (Knight, 1990)

PARAMETER VALUE

L Axial pad length 2.24 in. (5.7 cm)

tp Pad thickness 0.63 in. (1.6 cm)

Cp Radial pad clearance 4.72 mils (120 µm)

∆p Angular dimension of pad 54 deg

θp Pivot angles of pads 18, 90, 162, 234, 306 deg

m Preload factor 0.7

fp Fractional angular position of pivot 0.5

Rj Radius of journal 1.5 in. (3.8 cm)

ω Angular journal speed 9000 RPM (942 rad/s)

Wb Bearing load 2001 lbf (8900 N)

P0 Lubricant supply pressure 19.1 psi (0.132 MPa)

T0 Lubricant supply temperature 122 oF (50oC)

β Thermoviscosity index 0.017 oF-1 (0.031 oC-1)

µ0 Lubricant supply viscosity 1.74 µreyn (12 mPa s)

ρl Lubricant density 0.03 lbm/in3 (845 kg/m3)

cl Lubricant specific heat capacity 0.53 Btu/lbm/ oF (2200 J/kg/K)

The temperature at the leading edge of pad #5 was calculated for different mesh sizes

(NEX = number of elements in the x-direction). Since eq3. 52 is one dimensional, only

84
the θ or x direction meshing will be considered. The results from this analysis are shown

in Table 3. 7.

Table 3. 7 - Data Generated from Bearing Described in Table 3. 6

Temperature @
NEX leading edge of
Pad #5 (oF)
5 157

10 158

15 158

Table 3. 7 shows that the solution has converged for NEX = 10. The pressure calculation

also requires NEX to be at least 10 for the journal position to converge. Therefore, the

mesh size used will be NEX = 10 and NEZ = 6 which gives a total of 60 elements per pad.

The resulting temperature of 158 oF in Table 3. 7 is close to the actual experimental value

of about 156 oF, and gives a better approximation than Knight’s theoretical value of 131
o
F. The complete temperature profile for all of the pads is given in Figure 3. 9.

85
Figure 3. 9 - Temperature Distribution for Knight’s Bearing

The current model gives a reasonable approximation to the temperature distribution as

shown in Figure 3. 9. Nevertheless, this model has its limitations. It does not account for

heat transfer at the pad boundaries, and therefore does not predict the experimental

temperature drops at the trailing (end) edges of the pads. Knight’s theoretical model does

take into account this type of heat transfer and thus mimics these temperature drops as

indicated in Figure 3. 9. These temperature drops are important for a physically consistent

model. The TPJB is a circular system, which means that the trailing edge of pad 2 shares

a lubricant pocket with the leading edge of pad 3 and the temperature within that pocket

has to be some average of the pad 2 outlet and pad 3 inlet temperatures. Therefore, it is

necessary for the pad 2 trailing edge temperature to drop in order to avoid a physically

86
implausible temperature discontinuity. Even though the current model does not depict this

temperature drop, it does give a decent overall estimate for the temperature profile. In

fact, especially for pads 3 to 5, the current model gives a better approximation than

Knight’s theoretical model.

The temperature profile, the pressure distribution and the film thickness can also be

obtained at the bearing centerline (z = L/2) as shown in Figure 3. 10.

Figure 3. 10 - Film Thickness, Pressure and Temperature at the Centerline of


Knight’s Bearing (Table 3. 6)

As indicated in Figure 3. 10, the maximum temperatures usually occur on the most highly

loaded pads (greatest pressures) which are associated with the lowest film thickness

87
values. This is in agreement with physical expectations. The graphics program was

written in Matlab and is given in Appendix C.

3.5. Case Studies

3.5.1. de Jongh and Morton


A synchronous vibration problem was encountered by de Jongh and Morton (de

Jongh and Morton, 1994) in their analysis of a LP/IP centrifugal compressor. The

compressor was engaged in an off-shore gas-lift operation and its rotor was supported by

two 5-pad LOP TPJBs. It was found that it was impossible to attain the maximum

continuous operating speed (MCOS) of 11947 RPM because of unstable vibrations

encountered around 11400 RPM near the non-driven end (NDE) bearing. These fast-

growing vibrations reached a significant level and thus forced the rig operators to reduce

the rotor speed.

Upon examining the rig in a balancing facility it was discovered that the rotor was

very sensitive to small unbalances. It was also observed that the vibration problem

persisted when the mechanical and labyrinth seals were removed. This precluded the

possibility of the Newkirk Effect. However, the hysteresis in the vibration growth and

decay seemed to suggest a thermal phenomenon. De Jongh and Morton then built a test

rotor with identical dynamic characteristics as the full-sized compressor rotor and

measured the temperature difference across the shaft within the bearings. Resulting data

confirmed that the vibration problem was caused by thermally-induced bending and was

therefore a manifestation of the Morton Effect. The final solution to the vibration problem

was to reduce the overhang masses.

88
Using the data in deJongh and Morton paper, a VT-FAST model of the centrifugal

compressor was constructed. The compressor stages and other rotor portions were

simplified and modeled as disks, but the dimensions and gyroscopics were essentially the

same. Figure 3. 11 shows the resulting model and Table 3. 8 gives the corresponding

input data before any modifications were done to solve the thermal instability problem.

NDE DE

Figure 3. 11 - VT-FAST Model of deJongh and Morton Centrifugal Compressor

Rotor

89
Table 3. 8 - Input Data for deJongh and Morton Compressor Rotor Before

Modifications

PARAMETER NDE VALUE DE VALUE


Pad Parameters
L, Axial pad or bearing length 1.93 in. (4.9 cm) 1.93 in. (4.9 cm)
tp, Pad thickness 0.5 in. (1.27 cm) 0.5 in. (1.27 cm)
Cp, Radial pad clearance 5.8 mils (147 µm) 5.8 mils (147 µm)
∆p, Angular dimension of pad 56 deg 56 deg
θp, Pivot angles of pads 54,126,198,270,342 deg 54,126,198,270,342 deg
m, Preload factor 0.5 0.5
fp, Frac. ang. position of pivot 0.5 0.5

Journal/Bearing Parameters
Rj, Radius of journal 1.93 in. (4.9 cm) 1.93 in. (4.9 cm)
Wb, Bearing load 488 lbf (2172 N) 506 lbf (2250 N)
Md, Overhang Mass 59.2 lbm (26.9 kg) 74.0 lbm (33.6 kg)
Ld, Overhang Distance 7.0 in. (0.18 m) 4.7 in. (0.12 m)
ωMCOS, Max. Cont. Op. Spd. 11947 RPM (1251 rad/s) 11947 RPM (1251 rad/s)
Um, Initial Mechanical Unbalance 0.39 oz in. (0.28 kg mm) 0.39 oz in. (0.28 kg mm)

Lubricant Parameters
P0, Lubricant supply pressure 19.1 psi (0.132 MPa) 19.1 psi (0.132 MPa)
T0, Lubricant supply temperature 122 oF (50oC) 122 oF (50oC)
β, Thermoviscosity index 0.017 /oF (0.031/ oC) 0.017 /oF (0.031/ oC)
µ0, Lubricant supply viscosity 2.94 µreyn (20.3 cP) 2.94 µreyn (20.3 cP)
ρl, Lubricant density 0.03 lbm/in3 (860 kg/m3) 0.03 lbm/in3 (860 kg/m3)
cl, Lubricant sp. heat capacity 0.48 Btu/lbm/ oF 0.48 Btu/lbm/ oF
(2000 J/kg/K) (2000 J/kg/K)

In order to ensure that the model was similar to the deJongh and Morton compressor

rotor, the critical speed maps were compared as shown in Figure 3. 12. The undamped

critical speed maps show that there is reasonable agreement between the VT-FAST model

and the actual centrifugal compressor.

90
Figure 3. 12 - Critical Speed Maps for VT-FAST Model and Actual Compressor

The non-driven end (NDE) of the centrifugal compressor was then analyzed using the

Morton Effect program for the TPJB and the resulting unbalance curves are shown in

Figure 3. 13. These curves predict the onset of the instability at 11508 RPM which is very

close to the 11400 RPM threshold instability value observed by de Jongh and Morton.

91
Figure 3. 13 - Unbalance Curves for the Non-Driven End of the deJongh and

Morton Compressor (Before Modifications)

The hump around 10000 RPM in Figure 3. 13 is due to the 3rd critical speed (see

Figure 3. 12) which provides larger amplitude orbits. As mentioned in chapter 2, these

orbits reduce the film thickness near the hot spot and increase the temperature in this

region. More incentive for thermal bending is thus obtained and the unbalance level rises.

The response diminishes and hot spot film thickness increases after the rotor goes through

the critical. As a result, the thermal gradient and the overall unbalance decrease around

11000 RPM. A thermal instability is eventually achieved when the unbalance level rises

due to an increase in viscous dissipation at higher speeds. The 4th critical speed around

14500 RPM also contributes to this trend.

92
The driven-end (DE) of the compressor was also analyzed (Figure 3. 14) and this

end seemed to exhibit the Morton Effect at speeds greater than 12488 RPM.

Figure 3. 14 - Unbalance Curves for the Driven End of the de Jongh and Morton

Compressor (Before Modifications)

This DE instability was not mentioned in the paper by de Jongh and Morton (de Jongh and

Morton, 1994). However, their solution to the thermal instability at 11400 RPM involved

decreasing the overhang masses at both the NDE and the DE.

In order to stabilize the compressor rotor, de Jongh and Morton reduced the NDE

overhang mass by 27.6 lbm (12.5 kg) and decreased the DE overhang mass by by 17.9 lbm

(8.1 kg). These mass reductions were achieved primarily by reducing the density of the

overhang components -_ they replaced some of the steel components with titanium ones.

93
After these modifications, some of the model input data was changed as indicated in Table

3. 9. The input data not mentioned in Table 3. 9 is the same as that given in Table 3. 8.

Table 3. 9 - Input Data for de Jongh and Morton Compressor Rotor After

Modifications

PARAMETER NDE VALUE DE VALUE


Journal/Bearing Parameters
Wb, Bearing load 461 lbf (2051 N) 488 lbf (2171 N)
Md, Overhang Mass 31.6 lbm (14.3 kg) 56.1 lbm (25.4 kg)
Ld, Overhang Distance 4.8 in. (0.12 m) 4.0 in. (0.10 m)
Um, Initial Mechanical Unbalance 0.37 oz in. (0.27 kg mm) 0.37 oz in. (0.27 kg mm)

By using the data in Table 3. 9, unbalance curves were obtained for the NDE (Figure 3.

15). These curves show that the thermal instability is not present in the given speed range

and the MCOS of 11947 RPM can be safely achieved.

The reduction in the NDE overhang mass stabilizes the rotor system in several

ways with respect to the Morton Effect:

1. The lowered mass forces the 4th critical (at around 145000 RPM) to a higher speed

and thus reduces the influence of this critical on the lower speed orbits. As a result,

the resultant unbalance curves (U) are attenuated in magnitude.

2. Decreasing the overhang mass directly reduces md which, in turn, lowers the thermal

unbalance (Ut) and subsequently decreases U.

3. Since the mass reduction occurs primarily at the ends of the rotor, the center-of-

gravity of the overhung mass would be drawn closer to the NDE bearing and the Ld

value (overhang distance) would diminish (compare Table 3. 8 and Table 3. 9). This

94
decrease in Ld would lower the thermal deflection and thus reduce Ut. Hence, U

would be further reduced and prevented from exceeding the threshold value (Uthr).

4. The unbalance peak due to the critical near 10000 RPM has been significantly

attenuated. This attenuation may be partially due to the increase in phase angle

between the initial mechanical (Um) and thermal (Ut) unbalances. Before

modifications, this angle was 31o near 10000 RPM but after modifications this angle

increased to 61.8o near 10500 RPM (near the new position of the attenuated peak).

As a result, Um and Ut are more out of phase which decreases the overall unbalance,

U. In addition, Ut itself is lowered because of the previously-discussed reasons.

Figure 3. 15 - Unbalance Curves for the Non-Driven End of the de Jongh and

Morton Compressor (After Modifications)

95
Figure 3. 16 shows that the mass decrease at the DE also seems to have removed the

thermal instability from the operating speed region . Therefore, it appears that there are

no Morton Effect possibilities close to the MCOS of 11947 RPM (1251 rad/s). This is in

concordance with what de Jongh and Morton observed.

Figure 3. 16 - Unbalance Curves for the Driven End of the de Jongh and Morton

Compressor (After Modifications)

96
3.5.2. de Jongh and van der Hoeven
Another example of the Morton Effect operating in rotors supported by TPJBs

was provided by de Jongh and van der Hoeven (de Jongh and van der Hoeven, 1998).

They analyzed a synchronous instability in two identical pipeline compressors which were

being used in a Dutch gas station to transport natural gas. These 584 lbm (265 kg)

compressors operated between 5370 RPM and 9400 RPM and one is schematically shown

in Figure 3. 17.

DE NDE

Figure 3. 17 - VT-FAST Model of a Compressor Analyzed by de Jongh and van der

Hoeven

97
During operation, these compressors exhibited large vibration levels above 7200

RPM and the machines had to be shut down. Analysis confirmed that these vibrations

were synchronous in nature and that there was a hysteresis in the amplitudes. This means

that the magnitude of the deceleration vibrations were up to 4 times the magnitude of the

acceleration vibrations over the same speed range. The phase readings from these

responses were also erratic. These observations pointed to a thermal instability and the

labyrinth seals were checked to determine whether the Newkirk Effect was the cause of

the problem. However, these seals did not seem to be initiating a rub and further

examination of the shaft did not reveal any scuff marks that are normally associated with

shaft rubbing. It was finally concluded that the Morton Effect was the source of the

instability and steps were taken to remedy this problem.

The researchers found that increasing the bearing clearance slightly would solve

this problem. Increasing the bearing clearance causes a less centered journal orbit which

leads to more overall cooling of the hot spot and attenuates the thermal gradient. As a

result, the Morton Effect is mitigated and the associated unstable vibration levels would be

reduced. A more detailed explanation on the relationship of orbit position and the Morton

Effect is provided in chapter 2.

Even though increasing the clearance was a solution to this problem, de Jongh and

van der Hoeven decided to adopt a different approach. They found that the minimum

required increase in bearing clearance was only 0.03% which was not a very practical

design specification. As a result, they invented a heat sleeve barrier which reduced the

temperature gradient across the shaft and decreased the amount of thermal bending caused

by the Morton Effect.

98
The current model was applied to the de Jongh and van der Hoeven case and the

required input data is given in Table 3. 10.

Table 3. 10 - NDE Data for Compressors Analyzed by de Jongh and van der Hoeven

PARAMETER NDE VALUE


Pad Parameters
L, Axial pad or bearing length 2.0 in. (5.08 cm)
tp, Pad thickness 0.5 in. (1.27 cm)
Cp, Radial pad clearance 7.0 mils (178 µm)
∆p, Angular dimension of pad 56 deg
θp, Pivot angles of pads 54,126,198,270,342 deg
m, Preload factor 0.5
fp, Frac. ang. position of pivot 0.5

Journal/Bearing Parameters
Rj, Radius of journal 2.0 in. (5.08 cm)
Wb, Bearing load 490 lbf (2180 N)
Md, Overhang Mass 249 lbm (113 kg)
Ld, Overhang Distance 10.5 in. (0.27 m)
ωMCOS, Max. Cont. Op. Spd. 9400 RPM (984 rad/s)
Um, Initial Mechanical Unbalance 0.37 oz in. (0.27 kg mm)

Lubricant Parameters
P0, Lubricant supply pressure 19.1 psi (0.132 MPa)
T0, Lubricant supply temperature 122 oF (50oC)
β, Thermoviscosity index 0.017 /oF (0.031/ oC)
µ0, Lubricant supply viscosity 2.94 µreyn (20.3 cP)
ρl, Lubricant density 0.03 lbm/in3 (860 kg/m3)
cl, Lubricant sp. heat capacity 0.48 Btu/lbm/ oF
(2000 J/kg/K)

The overhang mass at driven-end (DE) of the compressors was much smaller than that of

the non-driven end (NDE) which indicates that the Morton Effect would be more likely to

occur at the NDE (as observed by de Jongh and van der Hoeven). As a result, only the

NDE will be analyzed for the Morton Effect.

99
Before any Morton Effect analysis is done, the current model must be checked to

ensure that it dynamically matches the actual compressor. One way of accomplishing this

task is to compare the critical speed map from the model with values from the compressor

(Figure 3. 18).

Figure 3. 18 - Critical Speed Maps for VT-FAST Model and an Actual Compressor

Analyzed by de Jongh and van der Hoeven

Figure 3. 18 does not show as good a match between the compressor and the model as

does Figure 3. 12. However, Figure 3. 18 indicates that there is a reasonable match in the

region of the actual support (primarily from bearings) characteristic for the compressor.

This match means that there is good agreement between the model and the compressor

100
within the operating region of the compressor. It is therefore safe to proceed with the

Morton Effect analysis.

The unbalance curves for the model (Figure 3. 17) were generated by using the

data given in Table 3. 10. These curves are depicted in Figure 3. 19.

Figure 3. 19 - Unbalance Curves for the Non-Driven End of a Compressor Analyzed

by de Jongh and van der Hoeven (Before Modifications)

The unbalance curves predict the onset of thermal instability at 7070 RPM and this agrees

fairly well with the 7200 RPM value recorded by de Jongh and van der Hoeven.

Increasing temperature rise due to enhanced viscous dissipation at higher speeds is

primarily responsible for the shape of the unbalance curve. Critical speeds also seem to

influence the unbalance curve, e.g. around 5000 RPM and around 9500 RPM.

101
De Jongh and van der Hoeven solved this thermal instability by installing a heat

sleeve barrier around the shaft portion within the bearing. This barrier trapped a layer of

air close to the shaft and insulated it from the hot lubricant. As a result, there was minimal

heat transfer to the shaft and the thermal gradient across the journal was reduced by about

85%. If this reduction in thermal gradient is included in the current model, the following

curves will be obtained.

Figure 3. 20 - Unbalance Curves for the Non-Driven End of a Compressor (with

Heat Sleeve Barrier) Analyzed by de Jongh and van der Hoeven

According to Figure 3. 20, the model predicts that the compressor would be stabilized if

the heat sleeve barrier were used. This statement agrees with the experimental

observations of de Jongh and van der Hoeven.

102
3.5.3. Overhung Compressor
A compressor company graciously provided some experimental data for an

overhung compressor which seemed to exhibit the symptoms of the Morton Effect. This

compressor was supported by two 4-pad TPJBs which had load-between-pads

configurations. Further details are summarized in Table 3. 11.

Table 3. 11 - Data for Overhung Compressor

PARAMETER NDE VALUE DE VALUE


Pad Parameters
L, Axial pad or bearing length 2.75 in. (6.99 cm) 1.625 in. (4.13 cm)
tp, Pad thickness 0.874 in. (2.22 cm) 0.812 in. (2.06 cm)
Cp, Radial pad clearance 6.95 mils (177 µm) 4.35 mils (110 µm)
∆p, Angular dimension of pad 75.5 deg 74.9 deg
θp, Pivot angles of pads 45,135,225,315 deg 45,135,225,315 deg
m, Preload factor 0.44604 0.29885
fp, Frac. ang. position of pivot 0.5 0.5

Journal/Bearing Parameters
Rj, Radius of journal 2.25 in. (5.72 cm) 2.00 in. (5.08 cm)
Wb, Bearing load 302.9 lbf (1347 N) 55.7 lbf (248 N)
Md, Overhang Mass 123.82 lbm (56.2 kg) 43.04 lbm (19.5 kg)
Ld, Overhang Distance 9.68 in. (0.25 m) 7.36 in. (0.19 m)
ωMCOS, Max. Cont. Op. Spd. 12075 RPM (1264 rad/s) 12075 RPM (1264 rad/s)
Um, Initial Mechanical Unbalance 0.1386 oz in. (0.1 kg mm) 0.1386 oz in. (0.1 kg mm)

Lubricant Parameters
P0, Lubricant supply pressure 25.0 psi (0.172 MPa) 25.0 psi (0.172 MPa)
T0, Lubricant supply temperature 140 oF (60oC) 140 oF (60oC)
β, Thermoviscosity index 0.01309 /oF (0.024/ oC) 0.01309 /oF (0.024/ oC)
µ0, Lubricant supply viscosity 1.7 µreyn (11.7 cP) 1.7 µreyn (11.7 cP)
ρl, Lubricant density 0.03414 lbm/in3 0.03414 lbm/in3
(945 kg/m3) (945 kg/m3)
cl, Lubricant sp. heat capacity 0.4685 Btu/lbm/ oF 0.4685 Btu/lbm/ oF
(1962 J/kg/K) (1962 J/kg/K)

103
The experimental data obtained from the compressor is shown in Figure 3. 21 and Figure

3. 22 and it represents a general accel-decel between 0 to 12000 RPM. Several smaller

accel-decels were done between 10000 RPM and 12000 RPM in attempt to surpass the

instability. Unfortunately, these were unsuccessful and were responsible for the loops seen

on the graphs within this speed range.

Figure 3. 21 - Synchronous Vibration Data obtained from an X-Probe at the NDE of

the Overhung Compressor

104
Figure 3. 22 - Synchronous Vibration Data obtained from an X-Probe at the DE of

the Overhung Compressor

The synchronous vibrations, the hystereses in magnitudes and the erratic phase responses

depicted in Figure 3. 21 and Figure 3. 22 indicate that the Morton Effect might be the

cause of the problem. By using forced response data, which was provided by the

compressor company, unbalance curves for the Morton Effect were obtained (Figure 3. 23

and Figure 3. 24).

105
Figure 3. 23 - Unbalance Curves for the NDE of the Overhung Compressor

106
Figure 3. 24 - Unbalance Curves for the DE of the Overhung Compressor

Figure 3. 23 and Figure 3. 24 show that the NDE is predicted to exhibit the Morton Effect

between 10630 RPM and 12640 RPM while the DE is predicted to go unstable after

12990 RPM. The experimental results indicate that the NDE went unstable near 12000

RPM which agrees fairly well with the predictions. On the other hand, the DE also went

unstable around 12000 RPM even though the analysis does not predict any DE instability

in this region.

According to the data, the DE response decreased in amplitude from about 8000

RPM to 12000 RPM. At this point the vibrations suddenly increased to unstable values.

It is possible that the NDE went unstable first and the DE was sensitive to the resultant

unbalance at the NDE. Different portions of a rotor exhibit different sensitivities to

107
unbalances at different locations. As a result of these different sensitivities, the high

unbalances from the Morton Effect can apply a centrifugal force to an overhung region of

the rotor and thus lead to unstable vibrations elsewhere. Such vibrations would explain

the sudden response at the driven end of the overhung compressor.

108
4. Design of Test Rotor

One of the goals of this research was to initiate experimental work on the Morton

Effect. In order to accomplish this goal, an attempt was made to design and build a test

rig which would hopefully exhibit the Morton Effect.

Since the Morton Effect is a thermal bending phenomenon, the measurement of the

transverse journal temperature difference was initially considered. So far, this

measurement has only been accomplished by deJongh and Morton (deJongh and Morton,

1994) who installed Pt-100 RTDs inside the journal portion within the bearing at a

distance of 1.3 mm from the journal surface. A “special slipringless transmitter” was then

used to transfer data from the rotating shaft to a stationary data acquisition system.

Details about this transmitting device appear to remain confidential.

An alternative way of accomplishing the temperature measurement would be to

install thermocouples within the shaft. However, even though thermocouple wire is

cheap, the extra machining on the shafts could prove to be costly and this would elevate

the already steep price of the 303 stainless steel shafts. In addition, any asymmetric

machining could further unbalance the rotor and lead to mechanical vibration problems.

109
The second problem would be to transfer the data from the rotating shaft to the

stationary data acquisition system. Traditionally, slipring devices are used for this purpose

but these devices are usually very costly (one quotation for a slipring system from

Michigan Scientific Corporation was in excess of $11,400). Slipring brushes are also

subjected to frictional heating at high shaft speeds and this could lead to discrepancies in

the temperature measurements. Attempts were also made to find temperature sensitive

paints, crystals or crayons that could detect the temperature difference which is typically

around 10oC. Unfortunately, these efforts proved to be unfruitful and the idea to measure

the transverse temperature difference was abandoned due to high cost and lack of

appropriate devices on the market. Therefore, no thermocouples or the associated data

acquisition equipment were purchased.

Even though the transverse temperature gradient will not be measured, the design

of a test rotor would still be beneficial. The Morton Effect has other symptoms such as

hysteresis in vibration magnitude plots, erratic phase graphs and synchronous vibration

responses which could still be identified with the available capabilities and budget.

However, before the test rig could be designed, it was necessary to establish several

design parameters which would restrict the design space and hopefully yield a good

design.

4.1. Design Parameters


In order to reduce the cost and effort involved in building a test rig, it was decided

to modify an existing rotor. This rotor and its associated data acquisition system was last

used by Xiaoqiang Wang and is schematically illustrated in Figure 4. 1 (Wang, 1997).

110
Fluid film TPJBs

1 - Thrust Pad Holder 5 - Oil Tank 9 - Computer


2 - Ball 6 - Control Cabinet 10 - Frequency Voltage Converter
3 - Steel Cover 7 - Oil Pump 11 - Key Phasor Cond. Circuit
4 - 10 HP Motor 8 - Proximeter Console 12 - 16 Channel Data Acq. System
13 - Motor Speed Controller

Figure 4. 1- Schematic Diagram of Existing Test Rig

The existing rotor was built upon a table that was 8ft. (2.4 m) long, 2 ft. (0.6 m) wide and

3 ft. (0.9 m) high when the safety cover was in place. At the non-driven end (NDE), a ball

and holder was used to eliminate the axial vibrations of the rotor shaft which was primarily

supported by two 5-pad LOP spherical pivot fluid film bearings. These bearings were

supplied by a oil system that consisted of a 10 gallon (38 liter) reservoir of Teresstic - 32

light turbine oil at a flow rate of about 1.85 GPM (7 LPM). A thermostat was used to

measure and control the oil temperature. The 303 stainless steel rotor shaft or journal was

about 43 in. (1.1 m) long with a 1 in. (2.5 cm) diameter, and a large test disk was mounted

on the shaft at the bearing midspan. A Kop-Flex KD2 disk coupling was used to attach

the test shaft to the drive shaft at the driven-end (DE) of the rotor. Two ball bearings

were used to support the drive shaft which was belted to a 10 HP (7.5 kW) motor. The

111
belt allowed the motor speed to be stepped up by a factor of 3 which translates to an

operating rotor speed range of 450 RPM (7.5 Hz) to about 10,000 RPM (167 Hz). Three

pairs of probes (from Bently Nevada) were used to record the horizontal and the vertical

vibrations of the shaft. These probes were referenced to a synchronous key phasor trigger

which was at the end of the drive shaft that fit into the Kopflex coupling. All of the

transducer probes were attached to a proximeter console that relayed the signals to the

data acquisition system and computer.

It was decided that most the data acquisition system should be maintained but the

final display should use an ADRE 8-channel data acquisition instrument instead of the 16

channel box. This choice was made because only 7 probes (7 channels) would be used

and the ADRE 108 DAI also allowed a variety of plots and tabular data to be obtained.

Such advantages were not offered by the existing 16 channel box.

The rotor should also be modified to attain the overhung configuration (Figure 4.

2) that is required for the Morton Effect.

Figure 4. 2- Schematic Diagram Showing the Modified Rotor Configuration for The

Morton Effect

112
In order to minimize cost, bearing power loss and unwanted pedestal vibrations, the size

and weight of the modified rotor was kept to a minimum. These criteria, and the given

existing rotor foundation, were the basis for selecting a shaft length of 36 in. (91 cm) and

a 1 in. (2.5 cm) thick, 3 in. (7.6 cm) diameter disk at the NDE overhang position. The

shaft diameter also needed to be small to minimize the weight, but large enough to allow a

thermal gradient to develop across the shaft. As a result, 2 shaft diameters of 1.0 in (2.5

cm) and 1.5 in (3.8 cm) were considered in the design process. The bearings were to be

custom-made by Rotating Machinery Technology and thus the bearing clearance was also

chosen to be an unknown factor that had to be determined from the analysis of the

designs.

The various designs were analyzed using the following design parameters:

(1) Attainable Instability Speed. The average speed at which the Morton Effect would

occur should be within the 450-10000 RPM operating range of the rig.

(2) Operation Away from the Critical Speeds (Ncr). If any of the designs have an

instability speed close to a lateral critical speed of the rotor, then the unstable

vibrations may be blamed on the critical response. Therefore, a good design should

have the Morton Effect occurring at a speed which is separated as far as possible from

the lateral criticals. On the other hand, criticals do provide the large amplitude orbits

(see chapters 2 and 3) which favor the Morton Effect. Hence, if the final design does

encounter the Morton Effect near a critical speed, and if the Morton Effect response

can be clearly separated from that of the critical, then this design would be acceptable.

(3) φ should be close to 0 . An optimum configuration for the Morton Effect is one in
o

which the thermal and the mechanical unbalances are exactly in phase. This

113
arrangement causes the resultant unbalance (U) to be maximized and increases the

possibility of U > Uthr which is the Morton Effect criterion.

(4) Centered Orbit. A more centered synchronous orbit allows one side of the journal to

be constantly heat while the diametrically opposite portion is constantly cooled.

Hence, the likelihood of a transverse journal temperature gradient is increased and the

chances that the Morton Effect will occur are enhanced.

(5) Adequate Orbit Amplitude. The orbit must be large enough to generate the viscous

energy required for the Morton Effect. Larger orbits tend to reduce the film thickness

which increases the shear strain rate of the fluid. As a result, the shear stresses are

higher and more viscous dissipation is produced. If the orbit is too large then the

journal could contact the pads and annihilate the babbitt. The optimum orbit,

therefore, should be restricted to a suitable operating range (0.6 - 0.8 Cp was chosen

for the design process).

4.2. Analysis of Test Rotor Designs


The various designs, which are described in Table 4. 1, were analyzed using the

parameters given in the previous section. As mentioned before, only the shaft diameter

and the diametral bearing clearance (2Cb) were chosen as design variables. An overhang

length of 10 in. (25.4 cm) was arbitrarily chosen (see Figure 4. 2).

114
Table 4. 1- Description of Rotor Designs

Design Shaft Diameter Diametral Bearing Clearance

S1.0C2.50 1.0 in. (2.5 cm) 2.50 mils (63.5 µm)

S1.0C3.00 1.0 in. (2.5 cm) 3.00 mils (76.2 µm)

S1.0C3.50 1.0 in. (2.5 cm) 3.50 mils (88.9 µm)

S1.0C4.00 1.0 in. (2.5 cm) 4.00 mils (102 µm)

S1.0C4.50 1.0 in. (2.5 cm) 4.50 mils (114 µm)

S1.5C3.00 1.5 in. (3.8 cm) 3.00 mils (76.2 µm)

S1.5C4.00 1.5 in. (3.8 cm) 4.00 mils (102 µm)

S1.5C5.25 1.5 in. (3.8 cm) 5.25 mils (133 µm)

S1.5C6.00 1.5 in. (3.8 cm) 6.00 mils (152 µm)

S1.5C7.00 1.5 in. (3.8 cm) 7.00 mils (178 µm)

The average instability speed (Nmort) for the onset of the Morton Effect was evaluated for

each of the designs along with the second and third lateral undamped critical speeds. The

results are displayed in the figure below (Figure 4. 3).

115
10000

9000

8000
Speed (RPM)

Nmort
7000 Ncr2
Ncr3

6000

5000

4000
S1.0C2.50

S1.0C3.00

S1.0C3.50

S1.0C4.00

S1.0C4.50

S1.5C3.00

S1.5C4.00

S1.5C5.25

S1.5C6.00

S1.5C7.00

Case

Figure 4. 3 - Average Morton Effect Instability Speeds and Critical Speeds for the

Designs

All of the Nmort values appear to be within the 450 - 10000 RPM operating range of the rig

but some of the designs, e.g. S1.0C4.50, appear to have Nmort values that are too close to

critical speeds. This proximity may present problems if the Morton Effect response cannot

be distinguished from the vibrations caused by the excitation of the critical speed.

116
The phase difference between the mechanical and the thermal unbalances for the

different designs is shown in Figure 4. 4.

120

100

80
| | (deg)

60

40

20

0
S1.0C2.50

S1.0C3.00

S1.0C3.50

S1.0C4.00

S1.0C4.50

S1.5C3.00

S1.5C4.00

S1.5C5.25

S1.5C6.00

S1.5C7.00

Case

Figure 4. 4 - |Angle| between Mechanical and Thermal Unbalances for the Different

Designs

From Figure 4. 4 it can be inferred that S1.5C3.00 would have the most out-of-phase

mechanical and thermal unbalances. Based on this criterion, the S1.5C3.00 case would be

the least likely to exhibit the Morton Effect. On the other hand, the S1.5C6.00 case has

117
the most in-phase mechanical and thermal unbalances and would therefore be a better

candidate for the Morton Effect.

In order to find out whether the design orbits are centered, the static eccentricity

ratio was considered (Figure 4. 5). Designs with lower eccentricity ratios have more

centered orbits which favor the Morton Effect. The attitude angles are not considered

because these angles are usually zero for tilting pad journal bearings.

0.400

0.350

0.300

0.250

0.200

0.150

0.100

0.050

0.000
S1.0C2.50

S1.0C3.00

S1.0C3.50

S1.0C4.00

S1.0C4.50

S1.5C3.00

S1.5C4.00

S1.5C5.25

S1.5C6.00

S1.5C7.00

Case

Figure 4. 5 - Static Eccentricity Ratio at Nmort for Each of the Designs

118
Figure 4. 6 shows the orbit size and its required limitations for the designs.

Designs such as S1.0C3.00 have too small orbits and therefore may not produce enough

viscous energy to fuel the Morton Effect. However, designs like S1.5C5.25 fall within the

amplitude boundaries and are more subject to the Morton Effect.

6.00

5.00

4.00
Radial Displacement (mils)

NDE + ecc

3.00 0.6 Cp
0.8 Cp

2.00

1.00

0.00
S1.0C2.50

S1.0C3.00

S1.0C3.50

S1.0C4.00

S1.0C4.50

S1.5C3.00

S1.5C4.00

S1.5C5.25

S1.5C6.00

S1.5C7.00

Case

Figure 4. 6 - Orbit Size at Nmort for the Various Designs

All of the data presented so far can be summarized and analyzed using Table 4. 2.

The design with the 1.5” diameter shaft and 5.25 mils diametral clearance is the only one

119
to satisfy all of the design parameters. Hence, this design will be chosen as the basis for

developing the experimental test rig.

Table 4. 2 - Summary of Design Analysis

Attainable Nmort away Suitable φ: Centered Orbit


Design Nmort from Ncrs φ = 0°±20° Orbit: Amplitude:
ε £ 0.3 0.6-0.8 Cp
S1.0C2.50 x x x

S1.0C3.00 x x x

S1.0C3.50 x x

S1.0C4.00 x

S1.0C4.50 x x

S1.5C3.00 x x

S1.5C4.00 x x x x

S1.5C5.25 x x x x x

S1.5C6.00 x x x x

S1.5C7.00 x x x

4.3. Test Rotor Components


The increase in shaft diameter from 1 in. (2.5 cm) to 1.5 in. (3.8 cm) made it

necessary for additional existing rig components to be modified or replaced. Some of the

new rig components that were changed or replaced are discussed in the following sections.

120
4.3.1. Test Shaft
The test shaft (Figure 4. 7) was to be made from 303 stainless steel which is

suitable for machining and possesses superior fatigue resistance when compared with

other stainless steels.

Notes
• All keyway dimensions should have
Test Shaft
tolerances of +0.0020,-0.0000 inches.
• The dimension tolerance on the shaft
length is not critical. 303 stainless steel
• Shaft surface finish ~ 16 micro-inches.

Figure 4. 7 - 303 Stainless Steel Shaft

Originally, it was thought that the previous shaft supplier for the Rotor Lab, i.e. Winfred

Berg, would be able to furnish the shaft without problems. Unfortunately, this was not the

case. After several weeks of negotiating with Winfred Berg , it was decided to look for

other suppliers on the online Thomas Regional Database (www.thomasregional.com).

Several shaft suppliers from this search were then contacted and eventually Staples

Research Corp. (7036 Tech Circle, Manassas, VA 20109, Tel.: (888) 765-7654, Fax:

121
(703) 335-1234) was selected as the shaft supplier. Two shafts were ordered at about

$235 per shaft and shaft 1 was found to have a diameter of 1.4985” whereas 1.4989” was

the measured diameter of shaft 2.

4.3.2. Bearings
Rotating Machinery Technology, Inc., a member of the Rotordynamics Affiliates

Group, supplied the 5-pad LOP tilting pad bearings. Both of these bearings had a bore of

1.5042” which meant that shaft 1 would have a bearing clearance of 5.7 mils and shaft 2

would have a clearance of 5.3 mils. Since shaft 2 was closer to the recommended 5.25

mils then this shaft was selected for future experimental work.

The bearing fits into the housing as shown in Figure 4. 8. Unfortunately, the

housing was initially designed for a 1.0” diameter shaft, and so the apertures in the end-

shields were too small to accommodate the new 1.5” diameter shaft. New end-shields had

to be fabricated from Lexan which is more ductile and less likely to crack than perspex.

All of the machining was done by the ME machine shop.

122
Figure 4. 8 - Bearing Housing Components

4.3.3. Coupling
Only one of the Kop-Flex coupling hubs needed to be replaced. The hub adjoining

the 1.0” drive shaft was still going to be used since the drive shaft would be unchanged.

However, the driven test shaft had a diameter of 1.5”, which meant that the other coupling

hub (which was designed for a 1.0” shaft) had to be changed and some balancing holes

had to be drilled. After some tedious negotiating with Motion Industries, Inc., which is a

distributor for Kop-Flex, it was agreed to purchase an entire coupling with the machine

work done as specified in Figure 4. 9. The final cost for the total package was in excess

of $600.

123
Figure 4. 9 - Schematic Diagram of the Kop-Flex Coupling Showing the Required

Modifications

4.3.4. Overhang Disk


Two designs were considered for the overhung disk: one which required a keyway

to be secured to the shaft, and one which did not. The disk without a keyway would

require a keyless shaft mount which typically requires considerable installation/removal

torque and the shaft may be damaged in the process. Furthermore, the external diameter

of a $51.56 keyless mount from Trantorque is 2.375”. This large diameter only leaves

about 0.3125” on the 3.0” diameter disk for balancing holes which may become necessary

124
– a large unbalance due to asymmetry in the disk or in the shaft mount could prevent the

Morton Effect speed from being attained.

If the disk with the keyway was chosen then there would have to be balancing

holes in order to counteract the unbalance imposed by the presence of a keyway.

However, the shaft would not be damaged and the disk could easily be installed and

replaced. In addition, no more funds would have to be spent on a shaft mount and there

would be more room on the disk for balancing holes. The disk with the keyway (Figure 4.

10) was chosen for the overhang disk design in light of these reasons.

Overhang Disk

Figure 4. 10 - Overhang Disk Design

125
4.3.5. Inboard Disk
During operation of the rig it was found that an inboard disk was required for

additional balancing of the rotor. Since this disk was to be placed in between the bearings

no keyway could be used. Instead, a small screw would be used to secure the disk to the

shaft. This method of attachment prevented the use of a very large disk because high

inertial forces could easily dislodge the disk. As a result, a thin 3” diameter disk was

fabricated (see Figure 4. 11).

Inboard Disk

Figure 4. 11 - Inboard Disk Design

This inboard disk was checked before and after the rotor was run at different speeds. No

shifting in the disk position was observed and the disk did assist in the overall balancing of

the rotor.

126
4.4. Analysis of Final Test Rotor Design
After the rotor components were assembled and all of the modifications were

made, the final test rotor design resembled the drawing in Figure 4. 12.

Figure 4. 12 - The Final Design for the Morton Effect Test Rotor

A comparison of Figure 4. 12 with Figure 4. 2 shows that some changes have been made.

The major alterations were a shortening of the bearing span from 21.5” to 21”, the

inclusion of an inboard disk and the repositioning of some of the probes. These changes

were done primarily to allow the rotor to be better balanced. The actual Morton Effect

rotor is depicted in Figure 4. 13.

127
overhang DE
NDE inboard coupling
disk brg hsg
brg hsg disk

probes probes 1 ft. probes


1h, 2v 3h, 4v ruler 5h, 6v

Figure 4. 13 - Picture of the Actual Morton Effect Test Rotor

The rest of the systems (lubrication, speed control and data acquisition) were described in

the Design Parameters section. However, a few minor changes had to be made. The oil

flow rate was decreased to 1 GPM in order reduce the leakage of oil from the center hole

in the end-shields. In addition, the oil inlet temperature was increased to 100oF to ensure

that the oil flowed smoothly through the supply pipes, the bearing housing and the

drainage pipes. Figure 4. 14 illustrates the Morton Effect rotor and its accompanying

systems.

128
master
ADRE 108 power
DAI

control rotor
box

computer with
ADRE Software

lubricant reservoir with


thermocouple and heater pump

Figure 4. 14 - Picture of the Morton Effect Rotor and Associated Systems

As a result of the modifications, the dynamics of the rotor were slightly altered

and the analyses (including the Morton Effect analysis) had to be repeated in greater

detail. These analyses were essential because they would give a final prediction for the

Morton Effect speed range of the rotor.

The first step in the analysis was to obtain a VT-FAST model by using the

dimensions and information in the given drawings. This model is shown in Figure 4. 15.

129
NDE DE

Figure 4. 15 - VT-FAST Model of The Morton Effect Rotor

Input data for the VT-FAST model and the other analyses is listed in Table 4. 3. The

rotor was originally designed to have the Morton Effect at the non-driven end. However,

Table 4. 3 shows that the driven end has a significant overhang and may also be

susceptible to this thermal bending phenomenon. Consequently, both ends of the rotor

will be analyzed for the Morton Effect.

130
Table 4. 3 - Data for Morton Effect Rotor

PARAMETER NDE VALUE DE VALUE


Pad Parameters
L, Axial pad or bearing length 0.55 in. (1.4 cm) 0.55 in. (1.4 cm)
tp, Pad thickness 0.6 in. (1.52 cm) 0.6 in. (1.52 cm)
Cp, Radial pad clearance 5.3 mils (135 µm) 5.3 mils (135 µm)
∆p, Angular dimension of pad 56 deg 56 deg
θp, Pivot angles of pads 54,126,198,270,342 deg 54,126,198,270,342 deg
m, Preload factor 0.5 0.5
fp, Frac. ang. position of pivot 0.5 0.5

Journal/Bearing Parameters
Rj, Radius of journal 0.75 in. (1.9 cm) 0.75 in. (1.9 cm)
Wb, Bearing load 12.58 lbf (56 N) 13.94 lbf (62 N)
Md, Overhang Mass 6.64 lbm (3.0 kg) 8.57 lbm (3.9 kg)
Ld, Overhang Distance 6.04 in. (0.15 m) 4.35 in. (0.11 m)
Um, Initial Mechanical Unbalance 0.022 oz in. 0.022 oz in.
(0.016 kg mm) (0.016 kg mm)

Lubricant Parameters
P0, Lubricant supply pressure 20 psi (0.138 MPa) 20 psi (0.138 MPa)
T0, Lubricant supply temperature 100 oF (38 oC) 100 oF (38 oC)
β, Thermoviscosity index 0.017 /oF (0.031/ oC) 0.017 /oF (0.031/ oC)
µ0, Lubricant supply viscosity 4.24 µreyn (29.2 cP) 4.24 µreyn (29.2 cP)
ρl, Lubricant density 0.03 lbm/in3 (860 kg/m3) 0.03 lbm/in3 (860 kg/m3)
cl, Lubricant sp. heat capacity 0.48 Btu/lbm/ oF 0.48 Btu/lbm/ oF
(2000 J/kg/K) (2000 J/kg/K)

One of the options in the VT-FAST program is a stability analysis. This feature calculates

the damped critical speeds of the rotor system and thus provides a better estimate of the

critical speeds than the undamped critical speed map. A knowledge of the location of the

critical speeds is essential for both the safe operation of the rig and the Morton Effect

model. By using this stability option, the first four criticals were obtained (see the

following figures.)

131
Figure 4. 16 - Predicted 1st Damped Critical Speed for the Morton Effect Rotor

132
Figure 4. 17 - Predicted 2nd Damped Critical Speed for the Morton Effect Rotor

133
Figure 4. 18 - Predicted 3rd Damped Critical Speed for the Morton Effect Rotor

134
Figure 4. 19 - Predicted 4th Damped Critical Speed for the Morton Effect Rotor

Figure 4. 16, Figure 4. 17, Figure 4. 18 and Figure 4. 19 show the predicted damped

critical speeds for the Morton Effect rotor. These critical speeds are around 4344 RPM,

5443 RPM, 8150 RPM and 24266 RPM respectively. Therefore, within the rotor’s

operating speed range of 450 RPM to 10000 RPM, there are 3 critical speeds. A forced

response analysis was done in order to determine which of these critical speeds would

produce unsafe vibration levels. In this analysis, an unbalance was applied to the rotor (at

the NDE overhang mass c.g.) and the resulting vibration levels were plotted as a function

of speed. This plot is shown in Figure 4. 20.

135
Figure 4. 20 - Response at Probes 3,4 due to a 1oz in. Unbalance at the NDE

Overhang Mass C.G.

From Figure 4. 20, it can be seen that the 3rd mode (near 8150 RPM) produces the most

significant response to the applied unbalance. The 1st and 2nd modes have either been

damped out or have not been sufficiently excited. Nevertheless, Figure 4. 20 provides a

guideline for what to expect when the rotor is running.

The final step in the analysis is to obtain the unbalance curves for the Morton

Effect. Both the NDE and the DE were analyzed and the resulting curves are shown in

Figure 4. 21 and Figure 4. 22 respectively.

136
Figure 4. 21 - Unbalance Curves for the Non-Driven End of the Morton Effect Test

Rotor (Final Design)

137
Figure 4. 22 - Unbalance Curves for the Driven End of the Morton Effect Test Rotor

(Final Design)

The figures above indicate that both ends of the rotor would be susceptible to the Morton

Effect. An instability range of 7280 RPM - 8380 RPM is observed for the NDE, whereas

the DE seems to undergo the Morton Effect from 7744 RPM - 8256 RPM. Besides

overlapping each other, these two ranges also encompass the 8150 RPM 3rd critical speed.

As previously-mentioned, critical speeds can produce large responses which can aid the

Morton Effect. However, there should be identifiable characteristics, e.g. hysteresis in

vibration amplitude and erratic phase angles, which can be used to distinguish the critical

speed behavior from the Morton Effect instability. It was therefore decided to run the

rotor in order to determine: (1) whether the Morton Effect would occur within the

138
predicted speed ranges and (2) if the thermal bending response could be distinguished

from that of the normal excitation of the 3rd critical speed.

139
5. Experimental Results from Test Rotor

After assembly (see chapter 4), the rotor had to be balanced in order to attain the

predicted Morton Effect speed range. Two types of experimental data were then

obtained by running the rotor close to and within the predicted instability region. The first

data set involved holding the rotor at a constant speed, for several minutes, within the

Morton Effect range. An acceleration and a deceleration through the Morton Effect speed

range produced the second set of data. Both of these data sets and the rotor balancing are

discussed in the following sections.

5.1. Rotor Balancing


All real rotors have some inherent mechanical unbalance which is caused by a mass

displaced from the axis of rotation. This mass produces a centrifugal force which can

drastically increase the rotor vibrations to levels that can damage the bearings. Therefore,

in order to safely attain the rotor operating speeds, it is frequently necessary to reduce

these unbalance levels to more acceptable values. This process is called rotor balancing.

The general purpose of rotor balancing is to reduce the current vibrations (As-Is

Vibrations) in the desired speed range to smaller values which are observed at lower

140
speeds (Run-Out Vibrations). Such a reduction is achieved by applying correction masses,

or trial weights, to the available balancing planes on the rotor.

The Morton Effect rotor was particularly difficult to balance. The primary

obstacle was to reduce the vibration levels near the 3rd critical speed (~ 8000 RPM) to

more acceptable values. It was originally thought that single-plane balancing would have

solved this problem. However, this technique proved to be ineffective. Two-plane

balancing was then tried with no avail. Finally, three-plane balancing was done and the

rotor was successfully run up to 9500 RPM. A least squares program (written by Dr.

Kirk) was used to calculate the correction masses for the 3 balancing planes. A

comparison of the unbalance levels before and after balancing is shown in Figure 5. 1.

Figure 5. 1 - Vibration Readings from Probe 5 (Before and After Balancing)

141
Figure 5. 1 shows that the response has been attenuated by about 1 mil and that the 3rd

critical was successfully traversed as the rotor was accelerated up to 9500 RPM. It should

be noted that, as predicted in chapter 4, the 1st and 2nd critical speeds did not present

serious vibration problems and were easily surpassed.

According to Vance (Vance, 1988), the difficulty encountered in balancing the

rotor may have been due to the balancing speed (7900 RPM) being too close to a bending

mode (3rd critical at ~ 8000 RPM). The elastic deformation of the rotor in this mode

redistributes the mass about the rotational axis. As a result, the unbalance configuration

can vary as this mode is approached and such behavior complicates the balancing process.

Another source of unbalance variation could have been the slight misalignment of

the test shaft. It was found that the coupling did not completely uncouple the two shafts

and, as a result, the orientation of the drive shaft skewed the test shaft position. Since the

coupling is flexible (but not enough to adequately uncouple the test and the drive shafts),

this degree of skew could vary as the rotor is brought up to speed. Such a variation

would cause a portion of the test shaft to be non-uniformly displaced from the rotation

axis and thus contribute to a varying mechanical unbalance. This behavior would present

an additional obstacle to balancing the rotor.

Nevertheless, these problems seemed to have been solved by the three-plane

balancing and the rotor was now suitable for experimental work.

5.2. Fixed Speed Data


Since the Morton Effect is a thermal phenomenon, it may require some time for the

thermal gradient to develop within the bearing. In the case of the turbocharger studied by

142
Faulkner, Strong and Kirk (Faulkner et al., 1997), it took about half an hour for this

phenomenon to occur at relatively constant speed. On the other hand, the Morton Effect

occurred after ~ 15 minutes at constant speed in the de Jongh and van der Hoeven

compressor (de Jongh and van der Hoeven, 1998). The current rig has a smaller shaft

diameter than either the turbocharger or the compressor and it may take less time for a

thermal gradient to develop in this shaft. However, smaller shaft diameters create less

viscous dissipation which would decrease the magnitude of the thermal gradient. In other

words, there was no certainty as to when the Morton Effect would occur in the current

rig.

In order to solve this dilemma, it was decided that the rotor should be held at a

fixed speed, for several minutes, within the predicted Morton Effect range. This process

would then be repeated for different speeds within the predicted instability region. A

typical plot from these experiments is shown in Figure 5. 2. This plot shows that the rotor

was taken up to 7800 RPM and held there for about 50 minutes. None of the probes,

including probe 3, recorded the significant increase in vibration levels that are normally

associated with the Morton Effect. The probe 3 response in Figure 5. 2 has some rippling

in its shape but this due to background electrical and mechanical noise. Therefore, the

fixed speed data did not yield any evidence of the Morton Effect.

143
8

0
14:45:36 14:52:48 15:00:00 15:07:12 15:14:24 15:21:36 15:28:48 15:36:00 15:43:12 15:50:24 15:57:36
Time (hh:mm:ss)

Probe 3 Response (mils p-p) Speed (kRPM)

Figure 5. 2 - Speed and Probe 3 Response for a Fixed Speed Data Plot

144
5.3. Accel-Decel Data
A second set of data was obtained by accelerating and decelerating the rotor

through a speed range which encompassed the predicted Morton Effect instability region.

Prior to this accel-decel, the oil was heated to 100oF at 500 RPM. The rotor speed was

then increased to 5000 RPM and data was recorded every 50 RPM until a speed of 9000

RPM was attained. At this point, the rotor was decelerated at the same rate (~ 4000

RPM/minute) back to 5000 RPM. Results from this accel-decel are shown in Figure 5. 3

and Figure 5. 4.

Figure 5. 3 - Probe 3 Response Data for 5000-9000 RPM Acceleration-Deceleration

(Oil Inlet Temperature = 100 oF)

145
Figure 5. 4 - Probe 5 Response Data for 5000-9000 RPM Acceleration-Deceleration

(Oil Inlet Temperature = 100 oF)

The figures above show the responses from a non-driven end (Probe 3) probe and a driven

end (Probe 5) probe. Both of these figures show responses that exhibit hysteresis in

amplitude and erratic phase. As the rotor is accelerated, both the DE and the NDE probes

show a peaking in amplitude around 8000 RPM and a phase drop that is expected when

critical speeds are traversed. However, during the deceleration, the amplitudes increase by

a factor of ~ 4 and the phase responses differ from the responses observed during

acceleration.

146
A cascade plot is a three-dimensional plot that has been collapsed into two

dimensions. It shows the magnitude and the frequency of the responses for different rotor

speeds. The cascade plots corresponding to the vibrations shown in Figure 5. 3 and

Figure 5. 4 are illustrated in Figure 5. 5 and Figure 5. 6 respectively.

Frequency (Hz)

Figure 5. 5 - Cascade Plot for Vibrations Shown in Figure 5. 3

147
Frequency (Hz)

Figure 5. 6 - Cascade Plot for Vibrations Shown in Figure 5. 4

The figures above indicate that the large deceleration vibration occurs at a frequency

which is equal to 1 X shaft rotational speed. Hence, this unstable response is synchronous

in nature.

By using the DC Gap voltage readings from the probes, the shaft centerline

position can be obtained as a function of speed. The NDE probes 3 and 4 were used to

obtain the location of the shaft center near the non-driven end bearing, while the DE

probes 5 and 6 were used to get the same data near the driven end bearing. This

information is displayed in Figure 5. 7 and Figure 5. 8.

148
Figure 5. 7 - Shaft Centerline Position (from Probes 3 and 4) near NDE Bearing

149
Figure 5. 8 - Shaft Centerline Position (from Probes 5 and 6) near DE Bearing

The NDE bearing (Figure 5. 7) appears to behave as a normal tilting pad bearing.

The attitude angle is close to zero and the shaft position rises with speed as a result of

increased lift. It should be noted that the 8000 RPM decel data point is the most centered

of all the shaft positions in Figure 5. 7.

Figure 5. 8 shows that the DE part of the test shaft is locked into a position close

to the bearing center and does not respond much to speed increases. This fixed position

may have been due to insufficient uncoupling between the drive shaft and the test shaft.

Such lack of uncoupling also presented problems during the balancing process.

In order to explain these interesting experimental results, several possibilities were

considered. These possibilities are discussed in the ensuing sections.

150
(1) Decrease in Support Damping. An rise in the decel response (see Figure 5. 3 and

Figure 5. 4) could have been caused by a decrease in damping within the bearing.

However, a decrease in damping would have also produced a steeper slope on the phase

curve at the ~ 8000 RPM critical speed. Figure 5. 3 and Figure 5. 4 show that this is not

the case. The slope of the phase curve at 8000 RPM seems to be the same for both the

accel and the decel cases. Hence, a decrease in damping was ruled out as a potential

explanation for the experimental results.

(2) Increase in Air Content of Lubricant. During the accel-decel runs, it was observed

that there was some foam present in the lubricant. These air bubbles in the oil could lead

to turbulent behavior which may produce the erratic results that were recorded. In order

to test this hypothesis, it was decided to increase the air content in the oil by lowering the

oil reservoir temperature. The colder oil would dissolve more air and should therefore

make the problem worse. Of course, lowering the oil temperature also increases the

viscosity and thus enhances the damping. Nevertheless, an instability of the recorded

magnitude should be able to supersede the increased damping if the driving force is the

increased air content in the lubricant. The results for the colder oil accel-decel runs are

shown in Figure 5. 9 and Figure 5. 10.

151
Figure 5. 9 - Probe 3 Response Data for 5000-9000 RPM Acceleration-Deceleration

(Oil Inlet Temperature = 78.3 oF)

152
Figure 5. 10 - Probe 5 Response Data for 5000-9000 RPM Acceleration-Deceleration

(Oil Inlet Temperature = 78.3 oF)

Figure 5. 9 and Figure 5. 10 indicate that the magnitude of instability is around 8 mils p-p

which is less that the ~ 11 mils p-p recorded in Figure 5. 3 and Figure 5. 4. Therefore, it

does not appear that the increased air content in the oil was the cause of the 8000 RPM

decel instability. However, more work should be done to verify this conclusion, e.g.

blowing air bubbles into the oil without significantly changing the temperature.

(3) Usual Accel-Decel Characteristic. Most rotors usually exhibit different response

characteristics in acceleration and deceleration. However, the deceleration response is

lower than the acceleration vibrations for both linear and non-linear systems (Ishida et al.,

1987). Figure 5. 3 and Figure 5. 4 show that the current rig displays the opposite trend,

153
i.e. the decel response is higher than the accel response. Therefore, this rotor does not

exhibit the normal accel-decel characteristics.

(4) Influence of Residual Energy from Critical. It was suggested during an interview with

Siemens Westinghouse Power Corporation that the deceleration might be influenced by

residual energy from the critical speed. Hence, the instability should be eliminated by

decelerating the rotor from a higher speed which is less influenced by the 8000 RPM

critical speed. The results from one such deceleration are shown in Figure 5. 11 and

Figure 5. 12.

Figure 5. 11 - Probe 3 Response Data for 5000-9500 RPM Acceleration-Deceleration

(Oil Inlet Temperature = 100 oF)

154
Figure 5. 12 - Probe 5 Response Data for 5000-9500 RPM Acceleration-Deceleration

(Oil Inlet Temperature = 100 oF)

The figures above show that the ~ 11 mils responses in Figure 5. 3 and Figure 5. 4 have

been reduced to about 7 mils. As a result, it does appear that the critical exerts some

influence over the decel instability. Operating further away from the critical reduces the

amount of energy put into the system at the start of the decel and thus reduces the overall

excitation of that critical. However, the instability still exists and the erratic phase shifts

are relatively unchanged. Nevertheless, the speed at which the decel begins does influence

the magnitude of the instability. More research needs to be done in this area, e.g. use a

more powerful motor to initiate a decel well above (> 10000 RPM) the 3rd critical speed

and observe the instability behavior.

155
(5) Increase in Mechanical Unbalance. A sudden increase in the mechanical unbalance

could also cause the high decel responses seen in Figure 5. 3 and Figure 5. 4. However,

such an elevation in the mechanical unbalance, e.g. blade loss in a turbine or loss of a

correction mass in the current rotor, would be present at all synchronous speeds and

would get worse at higher speeds. This type of behavior is not observed in the

experimental results. The rotor stabilizes after the instability and the magnitude decreases

at higher speeds. Furthermore, there is no significant synchronous vibration that is present

in the lower speed region. The experimental results are also repeatable which is contrary

to the effects of a sudden mechanical unbalance. Upon repeating the experiments, a

sudden unbalance would produce a more unstable system. A sudden mechanical

unbalance is therefore not responsible for the decel instability.

(6) Newkirk Effect. The hysteresis in the response, the erratic phase angles and the

synchronous vibrations can also indicate that a thermal instability, such as the Newkirk

Effect, could be present. However, there are no labyrinth (or any other) seals in the rotor

and there is a low probability of any contact between the rotor and a solid surface. The

most likely candidate to initiate a thermal rub would be the end shields and these have a

clearance of about 200 mils p-p. This value is much greater than the 11 mils p-p vibration

levels which were observed. Therefore, the Newkirk Effect may not be a potential source

of the recorded instability.

(7) Morton Effect. Another synchronous thermal instability, which could produce the

unstable response, would be the Morton Effect. The 8000 RPM instability peak is within

the predicted range for the Morton Effect (NDE: 7280 RPM - 8380 RPM, DE: 7744

RPM - 8256 RPM) and the experimental results also show that both ends go unstable

156
which is concordance with the analytical predictions. This synchronous instability also

shows a hysteresis in amplitude and yields erratic phase results. Furthermore, the shaft

centerline plots (Figure 5. 7 and Figure 5. 8) indicate that the orbits are fairly centered

near the unstable region. All of these characteristics describe or favor the Morton Effect.

Figure 5. 13 shows the response at the NDE bearing of the de Jongh and Morton

centrifugal compressor rotor which was discussed in chapter 3.

Figure 5. 13 - Synchronous Response at the NDE Bearing of de Jongh and Morton’s

LP/IP Compressor (de Jongh and Morton, 1994)

157
This rotor exhibited the Morton Effect and was studied intensively by de Jongh and

Morton (de Jongh and Morton, 1994). Figure 5. 13 indicates that the rotor was

accelerated up to about 11400 RPM when it encountered the Morton Effect. The rotor

was subsequently decelerated. However, since it takes some time for the thermal gradient

to subside, there is a hysteresis in the displacement plot and the decel amplitudes are

higher than the accel ones. In addition, the thermal bending causes the resultant unbalance

to vary in phase, i.e. the decel phase was higher than its accel counterpart. This erratic

phase variation may be due to the feedback nature of the Morton Effect. The initial

mechanical unbalance determines the response (phase and magnitude) at the bearing which

then leads to a thermal unbalance. The resultant unbalance can then cause a new response

at the bearing which can then invoke a new thermal unbalance. As a result of this process,

the phase of the response would change.

By comparing Figure 5. 13 with Figure 5. 3 or Figure 5. 4, one can see that there

are some similarities and some differences. Both sets of plots show that the decel

magnitude and phase data are generally higher than the corresponding accel values.

However, the de Jongh and Morton data show that the phase shifts and the amplitude

hystereses occur at the same speed, whereas Figure 5. 3 or Figure 5. 4 have the major

phase shifts occurring at lower speeds. In addition, the current rotor has the instability

occurring at a critical, but the de Jongh and Morton results have the Morton Effect in a

critical-free region. Even though critical speeds can aid the Morton Effect, they can also

make it difficult to distinguish the Morton Effect response from the vibrations due to the

excitation of the critical. Finally, the de Jongh and Morton compressor became

progressively unstable as the rotor was accelerated. The current rotor, on the other hand,

158
only became unstable when the rotor was decelerated. If the instability in the current rotor

was the Morton Effect then a possible explanation could be that the decel motion was

required to kick-start this phenomenon. Within the bearing lubricant, there is a velocity

gradient which causes the fluid closer to the shaft to be moving at higher velocities. As

the rotor is decelerated, some of this fluid retains the momentum from the original higher

speed and thus leads to an inertial effect at the current lower speed. It is possible that

such an inertial effect can alter the dynamics and reduce the phase difference between the

mechanical and thermal unbalances. This decreased phase separation would favor the

Morton Effect.

In summary, it can be said that the unusual instability observed may have been

caused by the Morton Effect or by influence from residual critical energy. More work

therefore needs to be done to investigate this instability. Besides extending the speed

range to check the influence of the critical, the temperature difference across the shaft

could be measured. This measurement would definitely support or refute the existence of

a thermal bending phenomenon. It should be noted, however, that such a measurement

would be both tedious and costly.

159
6. Conclusions

Several conclusions can be made based on the work presented in this dissertation.

These conclusions will be divided into those concerning the theoretical work and those

pertaining to the experimental research.

6.1. Conclusions based on the Theoretical Work


(1) Unlike most instabilities, which are non-synchronous, the Morton Effect is a

synchronous phenomenon. This synchronous thermal instability occurs primarily in

overhung rotors that are supported by fluid film bearings and is caused by differential

viscous shearing within the bearing lubricant.

(2) Most of the prior models of the Morton Effect consisted of frequency-domain

formulations with transient feedback characteristics.

(3) This dissertation proposes a new model to predict the onset of the Morton Effect. The

current steady-state model uses an unbalance threshold (based on 15% of the rotor

weight) as its instability criterion.

160
(4) The current model indicates that a worst-case scenario for the Morton Effect would be

one with a centered, circular and large amplitude orbit. Decreasing the phase difference

between the thermal and mechanical unbalances would also increase the likelihood of the

Morton Effect.

(5) Predictions made by the current model have both slightly underestimated and slightly

overestimated the actual instability speeds. However, there is generally good agreement

between the current model and other theoretical models and experimental data. This good

agreement applies to rotors which are supported by plain or tilting pad journal bearings.

6.2. Conclusions based on the Experimental


Research
(1) A test rotor was designed in order to initiate experimental work on the Morton Effect.

(2) During deceleration, this rotor exhibited an unusual instability near 8000 RPM. The

synchronous vibrations from this instability contained a hysteresis in magnitude and an

erratic phase response which are characteristics of the Morton Effect. Furthermore, this

instability occurred within the predicted Morton Effect speed range.

(3) The Morton Effect may have caused this unusual instability, but the influence of

residual energy from the 3rd critical speed (~8000 RPM) could also be the culprit.

161
7. Suggested Future Research

This dissertation represents an initial investigation of the Morton Effect at Virginia

Tech. It is hoped that it may provide some ground work for future studies. A few

suggestions and recommendations for future theoretical and experimental research are as

follows:

7.1. Suggested Future Theoretical Work

(1) More case studies can be done to test the program. Continued assistance from

companies who are willing to provide experimental data is encouraged.

(2) More investigations on the prediction of the onset of the Morton Effect as a function

of grid size could be done.

(3) Currently, the Morton Effect programs are designed to predict the onset of this

thermal bending phenomenon. These steady-state programs, however, do not give the

transient response that is available in other programs. Such a modification would allow

for better prediction of the magnitude and the phase of the thermal bending response.

162
7.2. Suggested Future Experimental Research

(1) Before any additional experimental work is done, it is suggested that the uncoupling

between the test shaft and the drive shaft should be increased. This task can be

accomplished by increasing the length of the coupling spacer tube. More uncoupling

would allow the DE bearing to behave more like a regular TPJB.

(2) As mentioned in chapter 5, more work can be done to investigate the effects of the

lubricant air content on the observed instability.

(3) The influence of the 3rd critical speed on the decel instability can also be checked by

operating at higher speeds. These experiments will require a more powerful motor to

supersede the current 10000 RPM maximum speed.

(4) Measurements of the transverse temperature gradient in the shaft portion (within the

bearing) could be done. This data would confirm or refute the existence of a thermal

bending phenomenon. However, such data would be difficult and expensive to obtain and

a new rig may have to be built. As a result, it is suggested that efforts can be made to

modify an existing rig (which exhibits the Morton Effect) from a willing company. The

resulting data can be shared with the company and a better understanding of the Morton

Effect (and of the validity of the programs) can be obtained.

163
Appendix A: Matlab GUI Program for Plain

Journal Bearings

%***********************************************************************
% clbck.m = program to execute when RUN button is pressed
%***********************************************************************

%Initialize program
clear all

%Percentage of rotor weight used in threshold unbalance


thrfac = 0.15;

%Number of averages for DT


NAVG = 10;

%Current figure handle


gfnum = gcf;

%Obtain inputs and convert to SI units

164
rhol = eval(get(findobj(gcbf,'Tag','rhol'),'String'))/(2.2046*0.0254^3);
capl = eval(get(findobj(gcbf,'Tag','capl'),'String'))*1055.1*2.2046*1.8;
mu0 = eval(get(findobj(gcbf,'Tag','mu0'),'String'))*6894.8;
beta = eval(get(findobj(gcbf,'Tag','beta'),'String'))*1.8;
T0 = (eval(get(findobj(gcbf,'Tag','T0'),'String'))-32)/1.8;
Tamb = (eval(get(findobj(gcbf,'Tag','Tamb'),'String'))-32)/1.8;
T02 = T0 - Tamb;

L = eval(get(findobj(gcbf,'Tag','L'),'String'))*0.0254;
Rj = eval(get(findobj(gcbf,'Tag','Rj'),'String'))*0.0254;
cb = eval(get(findobj(gcbf,'Tag','cb'),'String'))*1e-3*0.0254;
Wb = eval(get(findobj(gcbf,'Tag','Wb'),'String'))*4.4482;
alphaj = eval(get(findobj(gcbf,'Tag','alphaj'),'String'))*1.8;
HTC = eval(get(findobj(gcbf,'Tag','HTC'),'String'))*745.7*1.8/0.0254^2;

NS = eval(get(findobj(gcbf,'Tag','NS'),'String'));
W = eval(get(findobj(gcbf,'Tag','W'),'String'))*4.4482;
md = eval(get(findobj(gcbf,'Tag','md'),'String'))/2.2046;
Ld = eval(get(findobj(gcbf,'Tag','Ld'),'String'))*0.0254;
Um = eval(get(findobj(gcbf,'Tag','Um'),'String'))*0.0254/(2.2046*16);

speed = eval(get(findobj(gcbf,'Tag','speed'),'String'));
omega = speed*2*pi/60;
xamp = eval(get(findobj(gcbf,'Tag','xamp'),'String'))*1e-3*0.0254;
xph = eval(get(findobj(gcbf,'Tag','xph'),'String'))*pi/180;
yamp = eval(get(findobj(gcbf,'Tag','yamp'),'String'))*1e-3*0.0254;
yph = eval(get(findobj(gcbf,'Tag','yph'),'String'))*pi/180;

CaseText = get(findobj(gcbf,'Tag','case'),'String');

%Open output data file


fid = fopen(get(findobj(gcbf,'Tag','odf'),'String'),'w');
fprintf(fid,'%s\n\n\n',get(findobj(gcbf,'Tag','case'),'String'));

fprintf(fid,'%7s %7s %7s %7s %7s %7s %7s %7s\n','Speed',...


'Um','Ut','U','Uthr','DT','phi','eps0');
fprintf(fid,'%7s %7s %7s %7s %7s %7s %7s %7s\n\n','(RPM)',...
'(oz in)','(oz in)','(oz in)','(oz in)','(deg F)','(deg)',' ');

%Do calculations in speed loop


for i = 1:NS

%angular speed
w = omega(i);

%orbit parameters
Ax = xamp(i); Ay = yamp(i);
phx = xph(i); phy = yph(i);

%scale factor for the journal radius in the orbit plots


sf = 0.2*min([Ax Ay])/Rj;

%calculate heat transfer coefficient if not specified


if HTC > 0.0
H = HTC;
else
H = 0.5*25.5 * (w*Rj)^0.7 * mu0^-0.2 * (2*pi*Rj)^-0.4;
end

%calculate effective viscosity

165
A1 = 2*H/(rhol*capl*w*cb); B1 = 2*mu0*w*Rj^2/(rhol*capl*cb^2);
DTeff = (B1/A1-T02)*(1-exp(-A1*pi));
mu = mu0*exp(-beta*DTeff);

%epsilon = eccentricity ratio


a = (Wb*cb^2/(w*Rj*mu*L^3) )^2;

%get coordinates of journal static equilibrium position


eps = epsilon(a);
psi = atan(pi*sqrt(1-eps^2)/(4*eps));

thje = 1.5*pi + psi;


ecc = eps*cb;
xje = ecc*cos(thje);
yje = ecc*sin(thje);

%calculate angle which denotes hot spot location at t=0


g0 = atan2(Ay*sin(phy),Ax*cos(phx));

%parameters for temperature calculation


A = 2*H/(rhol*w*capl*cb);
B = 2*mu*w*Rj^2/(rhol*capl*cb^2);

%calculate average parameters for each orbit


for i2 = 1:NAVG
t = i2*2*pi/(w*NAVG);

%dynamic journal center position


xj = xje + Ax*cos(w*t + phx);
yj = yje + Ay*sin(w*t + phy);
thj = atan2(yj,xj);

%dynamic hot spot position


xh = xj + Rj*cos(w*t + g0);
yh = yj + Rj*sin(w*t + g0);
thh = atan2(yh,xh);
xih = mpippi(pi - (w*t + g0 - thj));

%dynamic cold spot position


xc = xj + Rj*cos(w*t + g0 + pi);
yc = yj + Rj*sin(w*t + g0 + pi);
thc = atan2(yc,xc);
xic = mpippi(pi - (w*t + g0 + pi - thj));

%solve differential equation for temperature distribution


%around journal circumference
eps2 = sqrt(xj^2+yj^2)/cb;
[th,T] = ode45('difeq2',[0 pi], [T02],[], A,B,eps2);

LT = length(T); T1 = T';
T2 = [T1(LT:-1:2) T1]; LT2 = length(T2);

nxih = round(1 + (LT2-1)*(xih+pi)/(2*pi) );


nxic = round(1 + (LT2-1)*(xic+pi)/(2*pi) );

if nxih == 0
nxih = 1;
elseif nxih > LT2
nxih = LT2;
end

166
if nxic == 0
nxic = 1;
elseif nxic > LT2
nxic = LT2;
end

%difference between hot spot and cold spot at


%various orbit positions
delT(i2) = (T2(nxih) - T2(nxic));

%compute angle used in resultant unbalance calculations


thch = atan2(yh-yc,xh-xc);
chi(i2) = w*t-thch;

%plot orbits
figure(gfnum+i)
orbplt(cb,delT(i2)*1.8,t,w,sf,Rj,Ax,Ay,phx,phy,xje,yje);

end

%calculate transverse temperature difference and


%thermal unbalance
DT = mean(delT);
yd = alphaj*DT*L*Ld/Rj;

Ut(i) = md*yd*2.2046*16/0.0254;
Utt = md*yd;

chiav = mean(chi);
phi = pi-chiav;

figure(gfnum+i)
title(sprintf('%s%s %7.0f %s %5.2f %s',CaseText,':',speed(i),...
' RPM, DT = ',1.8*abs(DT),' deg F'))
xlabel('x/Cb')
ylabel('y/Cb')
hold on

%total unbalance
U(i) = sqrt(Um^2 + Utt^2 - 2*Um*Utt*cos(chiav))*2.2046*16/0.0254;
Uthr(i) = (thrfac*W/w^2)*2.2046*16/0.0254;

%write data to file


fprintf(fid,'%7.0f %7.2f %7.2f %7.2f %7.2f %7.2f %7.1f %7.2f\n',...
speed(i),Um*2.2046*16/0.0254,Ut(i),U(i),Uthr(i),...
1.8*abs(DT),phi*180/pi,eps);

end

%Close data file


fclose(fid);

%Plot unbalance curves


figure(gfnum+NS+1)
sp2 = min(speed):0.001*(max(speed)-min(speed)):max(speed);
om2 = sp2*2*pi/60;
krpm = 1e-3*sp2;

unb = spline(speed,U,sp2);

167
unbthr2 = spline(speed,Uthr,sp2);

j = 1;
for i = 1:length(sp2)-1
diff1 = unb(i)-unbthr2(i);
diff2 = unb(i+1)-unbthr2(i+1);
if diff1*diff2 < 0
Nthr(j) = sp2(i);
Uval(j) = unb(i);
j = j + 1;
end
end

plot(krpm,unb,'b--'), hold on

%Threshold unbalance line


fac = thrfac;
unbthr = (fac*W./(om2.^2))*2.2046*16/0.0254;
plot(krpm,unbthr,'r-'), hold on

%Indicate threshold speeds if they exist


if j > 1
for i = 1:length(Nthr)
set(text(Nthr(i)*1e-3,Uval(i),...
sprintf('%7.0f %s',Nthr(i),'RPM')),'Rotation',[90],...
'Color',[0 0 0],'FontSize',[8])
plot(Nthr(i)*1e-3,Uval(i),'ko'), hold on
end
end

%Label plot
legend('U','Uthr')
title(CaseText)
xlabel('Speed (krpm)')
ylabel('Unbalance (oz in)')

%Display message box


if j > 1 | U(NS) > Uthr(NS)
msgbox(sprintf('%s%4.2f%s','Morton Effect will occur when U > Uthr
(Force = ',thrfac, 'W)'))
else
msgbox('Morton Effect will not occur within the given speed range')
end

%***********************************************************************

%***********************************************************************
% epsilon.m = function used by clbck.m. This function gives the static
% equilibrium eccentricity ratio for a narrow plain journal bearing.
%***********************************************************************

function [eps] = epsilon(a)

eps = real(sqrt(2 + sqrt((-3072*a + ...


(16 - pi^2 + ...
((-16 + pi^2)^3 + ...
72*a*(1024 - 64*pi^2 + 3*pi^4) + ...
12*sqrt(3)*sqrt(a)* ...
sqrt(67108864*a^2 + ...

168
pi^4*(-16 + pi^2)^3 + ...
4*a* ...
(-1048576 + 131072*pi^2 + ...
14336*pi^4 - 1152*pi^6 + ...
27*pi^8)))^0.3333333333333333)^ ...
2)/ ...
(a*((-16 + pi^2)^3 + ...
72*a*(1024 - 64*pi^2 + 3*pi^4) + ...
12*sqrt(3)*sqrt(a)* ...
sqrt(67108864*a^2 + ...
pi^4*(-16 + pi^2)^3 + ...
4*a* ...
(-1048576 + 131072*pi^2 + ...
14336*pi^4 - 1152*pi^6 + 27*pi^8)) ...
)^0.3333333333333333))/(4.*sqrt(3)) - ...
sqrt(384 - (4*(-16 + 96*a + pi^2))/a + ...
(3072*a - (-16 + pi^2)^2)/ ...
(a*((-16 + pi^2)^3 + ...
72*a*(1024 - 64*pi^2 + 3*pi^4) + ...
12*sqrt(3)*sqrt(a)* ...
sqrt(67108864*a^2 + ...
pi^4*(-16 + pi^2)^3 + ...
4*a* ...
(-1048576 + 131072*pi^2 + ...
14336*pi^4 - 1152*pi^6 + 27*pi^8) ...
))^0.3333333333333333) - ...
((-16 + pi^2)^3 + ...
72*a*(1024 - 64*pi^2 + 3*pi^4) + ...
12*sqrt(3)*sqrt(a)* ...
sqrt(67108864*a^2 + ...
pi^4*(-16 + pi^2)^3 + ...
4*a*(-1048576 + 131072*pi^2 + ...
14336*pi^4 - 1152*pi^6 + 27*pi^8))) ...
^0.3333333333333333/a - ...
(24*sqrt(3)*(-32 + pi^2))/ ...
(a*sqrt((-3072*a + ...
(16 - pi^2 + ...
((-16 + pi^2)^3 + ...
72*a*(1024 - 64*pi^2 + 3*pi^4) + ...
12*sqrt(3)*sqrt(a)* ...
sqrt(67108864*a^2 + ...
pi^4*(-16 + pi^2)^3 + ...
4*a* ...
(-1048576 + 131072*pi^2 + ...
14336*pi^4 - 1152*pi^6 + ...
27*pi^8)))^0.3333333333333333)^ ...
2)/ ...
(a*((-16 + pi^2)^3 + ...
72*a*(1024 - 64*pi^2 + 3*pi^4) + ...
12*sqrt(3)*sqrt(a)* ...
sqrt(67108864*a^2 + ...
pi^4*(-16 + pi^2)^3 + ...
4*a* ...
(-1048576 + 131072*pi^2 + ...
14336*pi^4 - 1152*pi^6 + ...
27*pi^8)))^0.3333333333333333))) ...
)/(4.*sqrt(3)))/sqrt(2));

%***********************************************************************

169
%***********************************************************************
% difeq2.m = function used by clbck.m. This function gives the
% differential equation for the circumferential temperature profile.
% The equation is solved by the Matlab ode45 command.
%***********************************************************************

function dT = difeq2(th,T,flag,A,B,eps)
dT = [-A*T(1)/(1+eps*cos(th)) + B/(1+eps*cos(th))^2];

%***********************************************************************

%***********************************************************************
% orbplt.m = function used by clbck.m. This function plots the total
% synchronous orbit and appropriately scales the journal radius to show
the
% locations of the cold and the hot spots at different points in the
orbit.
%***********************************************************************

function [] = orbplt(cb,dT,t,w,sf,Rj,Ax,Ay,phx,phy,xje,yje)

Rj = sf*Rj/cb;
Ax = Ax/cb; Ay = Ay/cb;
xje = xje/cb; yje = yje/cb;

g0 = atan2(Ay*sin(phy),Ax*cos(phx));
th = 0:0.01*pi:2*pi;

t1 = 0:0.01*2*pi/w:2*pi/w;
xp = xje + Ax*cos(w*t1 + phx);
yp = yje + Ay*sin(w*t1 + phy);

plot(xp,yp,'k-')
axis equal
hold on

xoj = xje + Ax*cos(w*t + phx);


yoj = yje + Ay*sin(w*t + phy);

xh = xoj + Rj*cos(w*t + g0);


yh = yoj + Rj*sin(w*t + g0);

xc = xoj + Rj*cos(w*t + g0 + pi);


yc = yoj + Rj*sin(w*t + g0 + pi);

xj = xoj + Rj*cos(th);
yj = yoj + Rj*sin(th);

plot(xh,yh,'rx',xc,yc,'bo',[xc xh],[yc yh],'g',...


xj,yj,'k')
axis equal

hold on

%***********************************************************************

%***********************************************************************
% mpippi.m = function used by clbck.m. This function places any given
% radian angle in the range -pi to +pi.
%***********************************************************************

170
function [xi2] = mpippi(xi)

if xi < -pi
while xi < -pi
xi = xi + 2*pi;
end

elseif xi > pi
while xi > pi
xi = xi - 2*pi;
end
end

xi2 = xi;

%***********************************************************************

171
Appendix B: Fortran Program for Tilting Pad

Journal Bearings
!*************************************************************************
! FINITE ELEMENT PROGRAM TO SIMULATE THE MORTON EFFECT OPERATING
! IN A CYLNDRICAL PIVOT TILTING PAD JOURNAL BEARING.
!*************************************************************************

!*************************************************************************
! PRE-PROCESSOR
!*************************************************************************
IMPLICIT NONE
INTEGER GDOF, EDOF, NBN, NBN2,NEX, NEZ, NE, NNX, NNZ, NPADS, NDOF,
* HBW, NPE, XXDOF, KCDOF, MAXITER, TGDOF, NSPD, NTAVG
REAL*8 PI, TOL, TOL2, DXX
PARAMETER ( NEX = 10, !# of elements in x direction
* NEZ = 6, !# of elements in z direction
* NE = NEX*NEZ, !total # of elements
* NDOF = 1, !number of nodal dof
* NPE = 4, !# of nodes per element
* EDOF = NDOF*NPE, !element degrees of freedom
* NNX = NEX + 1, !# of nodes in x direction
* NNZ = NEZ + 1, !# of nodes in z direction
* GDOF = NDOF*NNX*NNZ, !global degrees of freedom
* NBN = 2*NNZ + 2*(NNX-2),!# of press boundary nodes(EBCs)
* NBN2 = NNZ, !# of temp boundary nodes (EBCs)
* MAXITER = 20, !maximum # of iterations (Newton)
* NTAVG = 5, !# of trans.temp. averages
* DXX = 1D-6, !increment for XX
* TOL = 1D-6, !convergence tolerances
* TOL2 = 1D-6,
* HBW = (NNZ+2)*NDOF, !system half-bandwidth
* PI = 3.14159265359) !value for pi

REAL*8, DIMENSION(:), ALLOCATABLE::PRESS, TEMP, TEMP2,THTEMPC,


* TEMPC, PRESS2, PHI, SPEED, XAMP, XPH, YAMP, YPH,HIN,
* THETAP, EQDELTA, EQMOMENT, FX, FY, EQFX, EQFY, MOMENT,
* DEL, SOM, EQEPS, EQMU, EQPSI, HMIN, TMAX, PMAX,
* KXX, KYY, CXX, CYY, UNBT2, UNB2, UNBTHR2, PHIU,DELT
REAL*8 YD, LD, MD,UNBT, UNBM, UNB, UNBTHR, WROTOR
REAL*8 L, PADANGLE, M, OFFSET, THP, TP
REAL*8 RJ, RP, ECC, CB, CP, THC
REAL*8 N, U, MU, MU0, RHOL, CAPL, P0, T0, TIN, TAMB, DELTA
REAL*8 EQECC, EQTHC, EQSFX, EQSFY
REAL*8 X, Z, H, DHTH, BETA
REAL*8 SFX,SFY, LOAD, TIME
REAL*8 VX, VY, VD, RATIO, OMEGA, KIJ(4), CIJ(4)
REAL*8 ALPHAJ, DT, HTC, A, B, DTAV, DTAVG(NTAVG), SUMDTAVG
REAL*8 XJE, YJE, XJD, YJD, XHOT, YHOT, XCOLD, YCOLD,
* THCH, CHI(NTAVG),SUMCHI, CHIAV, GHOT, GCOLD, THHOT,THCOLD,
* TEMPHOT, TEMPCOLD

172
REAL*8 THS, MINHIN, TH
INTEGER PADNUM, PERTFLAG, SDFLAG, ITH1, ITH2
INTEGER I, I2, I3, IC, NPI, J, J2, J3, J4
CHARACTER:: INF1*20, OUTF1*20, OUTF2*20, TITLE*100

!------------------------------------------------------------------
!GET INPUT DATA FROM FILE
!------------------------------------------------------------------
PRINT *, 'Please enter the input file name:'
READ *, INF1

OPEN(1, FILE = INF1, STATUS = 'UNKNOWN')


READ(1,*)
READ(1,*) TITLE
READ(1,*)
READ(1,*)
READ(1,*)
READ(1,*)
READ(1,*) NPADS, L, TP, CP, PADANGLE, M, OFFSET

XXDOF = NPADS+2 !number of dof for Newton's Method


KCDOF = 4+5*NPADS !# of dof for complete Kij, Cij
TGDOF = NPADS*GDOF !GDOF for temperature dist.

ALLOCATE(PRESS(GDOF), TEMP(GDOF), TEMP2(TGDOF), TEMPC(NNX*NPADS),


* PRESS2(TGDOF), THETAP(NPADS), HIN(NPADS), THTEMPC(NNX*NPADS),
* EQDELTA(NPADS), EQMOMENT(NPADS), EQFX(NPADS), EQFY(NPADS),
* FX(NPADS), FY(NPADS), MOMENT(NPADS), DEL(NPADS))

READ(1,*)
READ(1,*)
READ(1,*) (THETAP(I), I = 1,NPADS)
READ(1,*)
READ(1,*)
READ(1,*)
READ(1,*)
READ(1,*) NSPD, RJ, ALPHAJ, LOAD

ALLOCATE(PHI(NSPD), SPEED(NSPD), XAMP(NSPD), XPH(NSPD),


* YAMP(NSPD), YPH(NSPD), SOM(NSPD), EQEPS(NSPD),EQMU(NSPD),
* EQPSI(NSPD), HMIN(NSPD), TMAX(NSPD), PMAX(NSPD),
* KXX(NSPD), KYY(NSPD), CXX(NSPD), CYY(NSPD),
* UNBT2(NSPD),UNB2(NSPD),UNBTHR2(NSPD),PHIU(NSPD),DELT(NSPD))

READ(1,*)
READ(1,*)
READ(1,*) LD, MD, WROTOR, UNBM
READ(1,*)
READ(1,*)

DO I = 1,NSPD
READ(1,*) SPEED(I), XAMP(I), XPH(I), YAMP(I), YPH(I)
END DO

READ(1,*)
READ(1,*)
READ(1,*)
READ(1,*)
READ(1,*) P0, T0, TAMB, MU0, BETA

173
READ(1,*)
READ(1,*)
READ(1,*) RHOL, CAPL
READ(1,*)
READ(1,*)
READ(1,*)
READ(1,*)
READ(1,*) SDFLAG ! SDFlAG = 0 => no coeff will be calculated
! SDFlAG = 1 => K&C coeff will be calculated
READ(1,*)
READ(1,*)
READ(1,*)
READ(1,*)
READ(1,*) OUTF1, OUTF2

CLOSE(1)

!------------------------------------------------------------------
!CONVERT INPUT DATA TO SI UNITS
!------------------------------------------------------------------

!Input Pad Parameters


L = L*0.0254D0 ! Axial bearing length (m)
TP = TP*0.0254D0 ! Pad thickness distance (m)
CP = CP*1.0D-3*0.0254D0 ! Radial pad clearance (m)
PADANGLE = PADANGLE*PI/180.0D0 ! Ang. dimension of pad (radians)
THETAP = THETAP*PI/180.0D0 ! Pad pivot angles (radians)

!Input Rotor Parameters


RJ = RJ*0.0254D0 ! Radius of journal (m)
ALPHAJ = ALPHAJ*1.8D0 ! Coeff. of thermal expansion (/K)
LOAD = LOAD*4.4482D0 ! Bearing load (N)
LD = LD*0.0254D0 ! Length to overhung mass (m)
MD = MD/2.2046D0 ! Overhung mass (kg)
UNBM = UNBM*0.02835*0.0254D0 ! Mechanical unbalance (oz in)
WROTOR = WROTOR*4.4482D0 ! Rotor weight (N)
XAMP = XAMP*1.0D-3*0.0254D0 ! Response: x-amplitude (m)
YAMP = YAMP*1.0D-3*0.0254D0 ! Response: y-amplitude (m)
XPH = XPH*PI/180.0D0 ! Response: x-phase (radians)
YPH = YPH*PI/180.0D0 ! Response: y-phase (radians)

!Input Fluid Parameters


P0 = P0*6894.8D0 ! Supply pressure to pad (Pa)
T0 = (T0-32.0)/1.8D0+273.15 ! Supply temperature to pad (K)
TAMB = (TAMB-32.0)/1.8D0+273.15 ! Ambient journal temperature (K)
MU0 = MU0*6894.8D0 ! Supply viscosity (Pa s)
BETA = BETA*1.8D0 ! Visc.-temperature parameter (/K)
RHOL = RHOL/(2.2046D0*0.0254**3) ! Density (kg/m^3)
CAPL = CAPL*1055.1*2.2046*1.8D0 ! Specific heat capacity(J/kg/K)

!Calculated Parameters
CB = CP*(1-M) ! Radial bearing clearance (m)
RP = RJ + CP ! Radius of curvature of pad

!Perturbation Parameters
RATIO = 0.005 ! Perturb. increment in ecc. ratio

!*************************************************************************
! PROCESSOR
!*************************************************************************

174
!Open unbalance data file
OPEN(10, FILE = OUTF2, STATUS = 'UNKNOWN')

!Begin speed loop


DO I2 = 1,NSPD

PRINT *, ' '


PRINT *, ' '
PRINT *, TITLE

N = SPEED(I2) ! Journal speed (RPM)


OMEGA = N*2*PI/60.0D0 ! Angular speed (rad/s)
U = OMEGA*RJ ! Journal speed (m/sec)
ECC = 0.4*CB ! Eccentricity (m) - initial guess
THC = 90.0*PI/180.0 ! Angle of line-of-centers(rads)
! - initial guess

HTC = 25.5 * U**0.7 * MU0**(-0.2) * (RJ*PADANGLE)**(-0.4)


! Brg heat transfer coeff. (W/m^2/K)

!------------------------------------------------------------------
! FIND EQUILIBRIUM POSITION
!------------------------------------------------------------------

! SPECIFY NO PERTURBATION
PERTFLAG = 0
VX = 0.0
VY = 0.0
VD = 0.0

! CALCULATE THE EFFECTIVE VISCOSITY


DTAV = -20.0
DO WHILE (DTAV <= 0.0)
A = 2.0*HTC/(1.0D0*RHOL*CAPL*OMEGA*CB)
B = 2.0*MU0*OMEGA*RJ**2/(1.0D0*RHOL*CAPL*CB**2)
DTAV = (B/(1.0D0*A)-(T0-TAMB))*(1-DEXP(-A*PI))
HTC = 0.9*HTC
END DO
MU = MU0*DEXP(-BETA*DTAV)

SOM(I2) = ( MU*(N/60.0D0)*L*(2*RJ)/(1.0D0*LOAD) )*(RJ/CP)**2


! Sommerfeld number

PRINT *, ' '


PRINT *, ' '
PRINT 110, 'SPEED = ',N, 'RPM'
PRINT 120, 'MU = ', MU*1.0D6/6894.8D0, 'microreyn'
PRINT 120, 'DTAV = ', DTAV*1.8, 'deg F'
PRINT 110, 'H = ',HTC*0.17612D0, 'Btu/(h ft^2 F)'
PRINT *, ' '
PRINT *, 'Calculating Equilibrium Position...'

!CALCULATE EQUILIBRIUM PRESSURE, TEMPERATURE AND POSITION


!Balance pad moments and X,Y forces on journal
CALL NEWTON(GDOF, EDOF, NBN, NEX, NEZ, NE, NNX, NNZ, NPADS,

175
* NDOF, HBW, NPE, XXDOF,MAXITER,PERTFLAG,TOL,
DXX,L,
* PADANGLE,M, OFFSET, THETAP, TP, RJ, RP, CB,
* P0, VX, VY, VD, LOAD,
* U, MU, SFX, SFY,FX, FY, DEL, MOMENT,ECC, THC)

PRINT *,' '


PRINT 120, 'EPS = ',DABS(ECC/CB), ' '
PRINT *,' '

PRINT *, 'Calculating Temperature Distribution...'


PRINT *,' '

DO I = 1,TGDOF
TEMP2(I) = T0
END DO

!Get inlet pad film thicknesses. These will help assign TIN
MINHIN = 10000.0

DO PADNUM = 1,NPADS

DELTA = DEL(PADNUM)
THP = THETAP(PADNUM)

THS = THP - OFFSET*PADANGLE

!Calculate the film thickness at each pad leading edge


CALL FILMTH(DELTA, M, THP,TP,RJ, ECC,CB,THC, RJ*THS, DHTH, H)

HIN(PADNUM) = H

IF (H < MINHIN) THEN


MINHIN = H
END IF

END DO

!Loop over pads


DO PADNUM = 1,NPADS

DELTA = DEL(PADNUM)
THP = THETAP(PADNUM)

!Calculate pad pressure distributions


CALL PRESSCALC(GDOF, EDOF, NBN, NEX, NEZ, NE, NNX, NNZ,
* NDOF, HBW, NPE, U, MU, P0, L, PADANGLE, M,OFFSET,
* THP, TP, VX, VY, VD, RJ, RP,ECC, CB, DELTA, THC, PRESS)

!Calculate the pad inlet temperature


TIN = T0 + 1.0D0*DTAV*MINHIN/HIN(PADNUM)

!Calculate pad temperature distributions


CALL TEMPCALC(GDOF, NEX, NEZ, NNX, NNZ,
* U, MU, RHOL, CAPL, T0, HTC, TIN, TAMB,
* L, PADANGLE, M,OFFSET,
* THP, TP, RJ, ECC, CB, DELTA, THC, TEMP)

!Store temperatures and pressures for each pad

176
DO I = 1,GDOF
J = GDOF*(PADNUM-1) + I
TEMP2(J) = TEMP(I)
PRESS2(J) = PRESS(I)
END DO

!End pad loop


END DO

!OBTAIN RELEVANT DATA AND WRITE THE PAD FILES


!Initialize critical output fluid parameters
HMIN(I2) = 100.0
PMAX(I2) = 0.0
TMAX(I2) = 0.0

!Extract equilibrium position data


DO PADNUM = 1,NPADS
EQDELTA(PADNUM) = DEL(PADNUM)
EQFX(PADNUM) = FX(PADNUM)
EQFY(PADNUM) = FY(PADNUM)
EQMOMENT(PADNUM) = MOMENT(PADNUM)
END DO

EQECC = ECC
EQTHC = THC
EQSFX = SFX
EQSFY = SFY

EQEPS(I2) = EQECC/(1.0D0*CB)
EQPSI(I2) = EQTHC*180.0D0/PI - 90.0
EQMU(I2) = MU

!Loop over pads to extract pressure and film thickness data


DO PADNUM = 1,NPADS
DELTA = DEL(PADNUM)
THP = THETAP(PADNUM)

SELECT CASE (PADNUM)

CASE(1)
OPEN(1, FILE = 'PAD1')
CASE(2)
OPEN(1, FILE = 'PAD2')
CASE(3)
OPEN(1, FILE = 'PAD3')
CASE(4)
OPEN(1, FILE = 'PAD4')
CASE(5)
OPEN(1, FILE = 'PAD5')
CASE(6)
OPEN(1, FILE = 'PAD6')
CASE(7)
OPEN(1, FILE = 'PAD7')
CASE(8)
OPEN(1, FILE = 'PAD8')
CASE(9)
OPEN(1, FILE = 'PAD9')
CASE(10)
OPEN(1, FILE = 'PAD10')

177
END SELECT

DO I = 1,GDOF

!Get global node coordinates


CALL NODCOORD(RJ, THP, NEX, NEZ, NNX, NNZ, OFFSET,
* PADANGLE, L, I, X, Z)

!Calculate the film thickness at each node


CALL FILMTH(DELTA, M, THP,TP,RJ, ECC,CB,THC, X, DHTH, H)

!Get the pressure and temperature at each node


J = GDOF*(PADNUM-1) + I
PRESS(I) = PRESS2(J)
TEMP(I) = TEMP2(J)

!Write data to appropriate file


!theta(deg), z(in.), T(deg F), P(psi)
WRITE (1,10) (X/RJ)*180/PI, Z/0.0254,
* 1.8*TEMP(I)-459.67,
PRESS(I)/6894.8,1.0D3*H/0.0254

!Extract critical parameters


IF(H < HMIN(I2)) THEN
HMIN(I2) = H
END IF

IF(PRESS(I) > PMAX(I2)) THEN


PMAX(I2) = PRESS(I)
END IF

IF(TEMP(I) > TMAX(I2)) THEN


TMAX(I2) = TEMP(I)
END IF

END DO

CLOSE (1)

!End pad loop


END DO

!CALCULATE STIFFNESS AND DAMPING COEFFICIENTS


IF (SDFLAG == 1) THEN

PRINT *, 'Calculating Stiffness and Damping Coefficients...'

CALL SDCOEFF(MAXITER,GDOF, EDOF, NBN, NEX, NEZ, NE, NNX,


* NNZ, NDOF, HBW, NPE, NPADS, U, MU, P0, L, PADANGLE, M,OFFSET,
* THETAP, TP, TOL2, DXX, VX, LOAD, VY, VD, RJ, RP, CB, OMEGA,
* EQECC, EQTHC, EQSFX, EQSFY, RATIO, PI, KIJ, CIJ)

KXX(I2) = KIJ(1)*0.0254/4.4482D0
KYY(I2) = KIJ(4)*0.0254/4.4482D0
CXX(I2) = CIJ(1)*0.0254/4.4482D0
CYY(I2) = CIJ(4)*0.0254/4.4482D0

END IF

178
!CALCULATE AVERAGE TRANSVERSE TEMPERATURE DIFFERENCE
IF (SDFLAG == 0) THEN

!Determine the journal equilibrium position


XJE = EQECC*DCOS(PI-EQTHC)*1.0D0
YJE = -EQECC*DSIN(PI-EQTHC)*1.0D0

!Angles to help define hot and cold spots


GHOT = DATAN2( YAMP(I2)*DSIN(YPH(I2)),XAMP(I2)*DCOS(XPH(I2)) )
GCOLD = GHOT + PI

IC = NINT(0.5*NNZ) !Centerline temperature


index
NPI = NINT(180.0*NPADS*NEX/360.0) !Node separation for 180 deg

!Write data to OUTF2


WRITE(10,240) XAMP(I2), XPH(I2), YAMP(I2), YPH(I2), XJE, YJE

!Begin DT averaging loop


DO J3 = 1,NTAVG

TIME = (J3-1)*2.0D0*PI/(1.0D0*OMEGA*NTAVG)

!Dynamic journal center postion


XJD = XJE + XAMP(I2)*DCOS(OMEGA*TIME + XPH(I2))
YJD = YJE + YAMP(I2)*DSIN(OMEGA*TIME + YPH(I2))

!Hot spot parameters


XHOT = XJD + RJ*DCOS(OMEGA*TIME + GHOT)
YHOT = YJD + RJ*DSIN(OMEGA*TIME + GHOT)
THHOT = DATAN2(YHOT,XHOT)

IF (THHOT < -PI) THEN


DO WHILE (THHOT < -PI)
THHOT = THHOT + 2*PI
END DO
ELSEIF (THHOT > PI) THEN
DO WHILE (THHOT > PI)
THHOT = THHOT - 2*PI
END DO
END IF

!Cold spot parameters


XCOLD = XJD + RJ*DCOS(OMEGA*TIME + GCOLD)
YCOLD = YJD + RJ*DSIN(OMEGA*TIME + GCOLD)
THCOLD = DATAN2(YCOLD,XCOLD)

IF (THCOLD < -PI) THEN


DO WHILE (THCOLD < -PI)
THCOLD = THCOLD + 2*PI
END DO
ELSEIF (THCOLD > PI) THEN
DO WHILE (THCOLD > PI)
THCOLD = THCOLD - 2*PI
END DO
END IF

!Chi = Angle for use in unbalance calculations


THCH = DATAN2(YHOT-YCOLD,XHOT-XCOLD)

179
CHI(J3) = OMEGA*TIME-THCH

!Update eccentricity and line of centers


ECC = DSQRT(XJD**2.0D0 + YJD**2.0D0)
THC = DATAN2(YJD,XJD) - PI

DO I = 1,TGDOF
TEMP2(I) = T0
END DO

PERTFLAG = 5

PRINT *, ' '


PRINT *, 'ORBIT AVERAGE # ',J3,' OF ',NTAVG

!Balance pad moments and X,Y forces on journal


CALL NEWTON(GDOF, EDOF, NBN, NEX, NEZ, NE, NNX, NNZ, NPADS,
* NDOF, HBW, NPE, NPADS,MAXITER,PERTFLAG,TOL2,DXX,
* L,PADANGLE,M, OFFSET, THETAP, TP, RJ, RP, CB,
* P0, VX, VY, VD, LOAD,
* U, MU, SFX, SFY,FX, FY, DEL, MOMENT,ECC, THC)

!Get inlet pad film thicknesses. These will help assign TIN
MINHIN = 1000.0

DO PADNUM = 1,NPADS

DELTA = DEL(PADNUM)
THP = THETAP(PADNUM)

THS = THP - OFFSET*PADANGLE

CALL FILMTH(DELTA, M, THP,TP,RJ, ECC,CB,THC, RJ*THS, DHTH, H)

HIN(PADNUM) = H

IF (H < MINHIN) THEN


MINHIN = H
END IF

END DO

!Loop over pads


DO PADNUM = 1,NPADS

DELTA = DEL(PADNUM)
THP = THETAP(PADNUM)

!Calculate the pad inlet temperature


TIN = T0 + 1.0D0*DTAV*MINHIN/HIN(PADNUM)

!Calculate pad temperature distributions


CALL TEMPCALC(GDOF, NEX, NEZ, NNX, NNZ,
* U, MU, RHOL, CAPL, T0, HTC, TIN, TAMB,
* L, PADANGLE, M,OFFSET,
* THP, TP, RJ, ECC, CB, DELTA, THC, TEMP)

!Store temperatures for each pad

180
DO I = 1,GDOF
J = GDOF*(PADNUM-1) + I
TEMP2(J) = TEMP(I)
END DO

!End pad loop


END DO

!Obtain centerline temperatures and corresponding angles


DO PADNUM = 1,NPADS

THP = THETAP(PADNUM)
THS = THP - OFFSET*PADANGLE

DO I = 1,NNX
I3 = IC + (I-1)*NNZ

J = GDOF*(PADNUM-1) + I3
J2 = NNX*(PADNUM-1) + I

TEMPC(J2) = TEMP2(J)

TH = THS + (I-1)*PADANGLE/(NEX*1.0D0)
THTEMPC(J2) = DATAN2(DSIN(TH),DCOS(TH))

END DO
END DO

!Compute difference between cold and hot spot temperature


DO J4 = 1,NPADS*NNX

IF (J4 == NPADS*NNX) THEN


ITH1 = NPADS*NNX
ITH2 = 1
ELSE
ITH1 = J4
ITH2 = J4+1
END IF

IF (THHOT >= THTEMPC(ITH1) .AND. THHOT <= THTEMPC(ITH2)) THEN


TEMPHOT = 0.5D0*(TEMPC(ITH1) + TEMPC(ITH2))
END IF

IF (THCOLD >= THTEMPC(ITH1).AND.THCOLD <= THTEMPC(ITH2)) THEN


TEMPCOLD = 0.5D0*(TEMPC(ITH1) + TEMPC(ITH2))
END IF

END DO

DTAVG(J3) = TEMPHOT - TEMPCOLD

!Write data to OUTF2


WRITE(10,220) DABS(DTAVG(J3))*1.8, 0.0, 0.0, 0.0, 0.0, 0.0

!End DT averaging loop


END DO

!Calculate averaged transverse temperature across journal

181
SUMDTAVG = 0.0D0
DO J3 = 1,NTAVG
SUMDTAVG = SUMDTAVG + DTAVG(J3)
END DO

DT = SUMDTAVG/(1.0D0*NTAVG)

PRINT *, ' '


PRINT 120, 'DT = ',1.8*DABS(DT),' deg F'

!Calculate averaged chi - angle for resultant unbalance calculations


!and obtain phi = angle b/w Ut and Um
SUMCHI = 0.0D0
DO J3 = 1,NTAVG
SUMCHI = SUMCHI + CHI(J3)
END DO

CHIAV = SUMCHI/(1.0D0*NTAVG)
PHI(I2) = PI-CHIAV

!Compute thermal deflection


YD = 1.0D0*ALPHAJ*DT*L*LD / RJ

!Calculate thermal unbalance


UNBT = MD*YD

!Calculate total unbalance


UNB = DSQRT(UNBT**2 + UNBM**2 - 2.0D0*UNBT*UNBM*DCOS(CHIAV))

!Calculate threshold unbalance


UNBTHR = 0.15D0*WROTOR/(1.0D0*OMEGA**2)

!Calculate phase of U
PHIU(I2) = DACOS((UNB**2 + UNBM**2 - UNBT**2)/(2.0D0*UNB*UNBM))

!Write data to OUTF2


WRITE (10,80) N, UNBTHR*2.2046*16.0D0/0.0254D0,
* UNB*2.2046*16.0D0/0.0254D0, DT*1.8, PHI(I2)*180.0D0/PI,
* EQECC/(1.0D0*CB)

UNBT2(I2) = DABS(UNBT*2.2046*16.0D0/0.0254D0)
UNB2(I2) = DABS(UNB*2.2046*16.0D0/0.0254D0)
UNBTHR2(I2) = DABS(UNBTHR*2.2046*16.0D0/0.0254D0)
DELT(I2) = DABS(DT*1.8)

END IF

!End speed loop


END DO

!Write data to OUTF2


DO I = 1,NPADS
WRITE(10,220) THETAP(I)*180.0D0/PI,0.0,0.0,0.0,0.0,0.0
END DO

182
WRITE(10,230) NTAVG, RJ, CB, WROTOR, 0.0, 0.0

WRITE(10,210) NPADS, NSPD, NNX, NNZ, L/0.0254D0,


* 1.8*TMAX(NSPD)-459.67

!Close unbalance data file


CLOSE (10)

!*************************************************************************
! POST-PROCESSOR
!*************************************************************************
!==================================================================
!WRITE INPUT/OUTPUT FILE
!==================================================================
OPEN (3, FILE = OUTF1, STATUS = 'UNKNOWN')

WRITE (3,*) TITLE

WRITE(3,*) ' '


WRITE (3,*) 'Pad Parameters'
WRITE (3,*) '--------------------'
WRITE (3,145) 'NPADS, Number of pads =',NPADS,' '
WRITE (3,20) 'L, Axial pad length =',L/0.0254,'inches'
WRITE (3,20) 'TP, Pad thickness =',TP/0.0254,'inches'
WRITE (3,20) 'CP, Radial pad clearance =', 1D3*CP/0.0254, 'mils'
WRITE (3,20) 'PADANGLE, Angular pad dimension =',
*
PADANGLE*180.0/PI,'deg'
WRITE(3,20) 'THETAP, Pivot angles =', THETAP(1)*180.0/PI,'deg'

DO I = 2,NPADS
WRITE(3,20) '=', THETAP(I)*180.0/PI,'deg'
END DO

WRITE (3,20) 'M, Preload factor =',M,' '


WRITE (3,20) 'OFFSET, Dimensionless pivot position =',OFFSET,' '
WRITE (3,*) ' '

WRITE (3,*) 'Rotor Parameters'


WRITE (3,*) '--------------------'
WRITE (3,20) 'RJ, Journal radius =', RJ/0.0254,'inches'
WRITE (3,20) 'CB, Radial bearing clearance =',1D3*CB/0.0254,'mils'
WRITE (3,20) 'ALPHAJ, Coeff. of thermal exp. =', ALPHAJ*1D6/1.8,
*
'/MF'
WRITE (3,20) 'LOAD, Bearing load =', LOAD/4.4482, 'lbf'
WRITE (3,20) 'WROTOR, Rotor weight =', WROTOR/4.4482, 'lbf'
WRITE (3,20) 'MD, Overhang mass =', MD*2.2046,'lbm'
WRITE (3,20) 'LD, Length to overhang mass =', LD/0.0254,'inches'
WRITE (3,20) 'UNBM, Mechanical overhang unbalance =',
* UNBM/(0.02835*0.0254D0),'oz in'
WRITE (3,*) ' '

WRITE (3,*) 'Lubricant Parameters'


WRITE (3,*) '--------------------'
WRITE (3,20) 'P0, Supply pressure =', P0/6894.8,'psi'
WRITE (3,20) 'T0, Supply temperature =',1.8*T0-459.67,'F'
WRITE (3,20) 'MU0, Supply viscosity =', 1D6*MU0/6894.8,'microreyn'
WRITE (3,200) 'BETA, Thermoviscosity coefficient =',
* BETA/1.8D0,'/F'

183
WRITE (3,200) 'RHOL, Density =', RHOL*2.2046*0.0254**3, 'lbm/in^3'
WRITE (3,200) 'CAPL, Specific heat capacity =',
* CAPL/4186.8,'Btu/(lbm F)'
WRITE(3,*) ' '

WRITE (3,*) 'Equilibrium Data'


WRITE (3,*) '--------------------'
WRITE (3,160) 'Speed','Sommerf','Ecc','Visc',
* 'HMIN','PMAX','TMAX'
WRITE(3,160) '(RPM)','Number','Ratio','microreyn',
* '(mils)','(psi)','(F)'

DO I = 1,NSPD
WRITE(3,170) SPEED(I),SOM(I),DABS(EQEPS(I)),
* EQMU(I)*1.0D6/6894.8,
* 1.0D3*HMIN(I)/0.0254,PMAX(I)/6894.8,1.8*TMAX(I)-459.67
END DO

WRITE (3,*) ' '

IF (SDFLAG == 1) THEN

WRITE (3,*) 'Bearing Stiffness and Damping Coefficients'


WRITE (3,*) '-------------------------------------------'

WRITE (3,180) 'Speed','Kxx','Kyy','Cxx', 'Cyy'


WRITE (3,180) '(RPM)','(lbf/in)','(lbf/in)',
* '(lbf s/in)','(lbf s/in)'

DO I = 1,NSPD
WRITE(3,190) SPEED(I),KXX(I),KYY(I),CXX(I),CYY(I)
END DO

END IF

WRITE (3,*) ' '

IF (SDFLAG == 0) THEN

WRITE (3,*) 'Unbalance Data'


WRITE (3,*) '--------------------'

WRITE (3,250) 'N','Um','Ut','U','Uthr','phi','DT'


WRITE (3,250) '(RPM)','(oz in)','(oz in)','(oz in)','(oz in)',
* '(deg)', '(deg F)'

DO I = 1,NSPD
WRITE (3,260) SPEED(I), UNBM/(0.02835*0.0254D0),
* UNBT2(I), UNB2(I), UNBTHR2(I), PHI(I)*180.0D0/PI,
* DELT(I)
END DO

END IF

CLOSE(3)

184
!*************************************************************************
! FORMATS
!*************************************************************************
10 FORMAT (5(E10.3,1X))
20 FORMAT (A40,1X,F8.2,1X,A16)
30 FORMAT (A6,1X,A11,1X,A8,1X,A8,1X,A8,1X,A11,1X,A16)
40 FORMAT (I2,5X,F5.0,7X,F8.0,1X,F8.0,1X,F8.0,1X,F7.5,5X,F7.3)
50 FORMAT (A40,1X,F8.4,1X,A10)
60 FORMAT (A40,1X,F8.0,1X,A10)
70 FORMAT (A40,1X,F8.1,1X,A10)
80 FORMAT (F6.0,1X,F6.2,1X,F6.2,1X,F6.2,1X,F6.2,1X,F6.2)
90 FORMAT (A4,1X,A10,1X,A15,1X,A4,1X,A10,1X,A15)
100 FORMAT (A4,1X,E10.4,1X,F10.2,6X,A4,1X,E10.4,1X,F10.2)
110 FORMAT (A10,1X,F10.0,1X,A15)
120 FORMAT (A10,1X,F10.2,1X,A15)
130 FORMAT (F10.0,2X,4(E10.4,2X))
140 FORMAT (A40,1X,F8.0,1X,A16)
145 FORMAT (A40,1X,I8,1X,A16)
150 FORMAT ('1',A4)
160 FORMAT (7(A8,1X))
170 FORMAT (F8.0,1X,F8.2,1X,F8.2,1X,F8.2,
* 1X,F8.1,1X,F8.0,1X,F8.0)
180 FORMAT (5(A10,1X))
190 FORMAT (F10.0,1X,4(E10.4,1X))
200 FORMAT (A40,1X,F8.3,1X,A16)
210 FORMAT (I6,1X,I6,1X,I6,1X,I6,1X,F6.2,1X,F6.0)
220 FORMAT (F6.2,1X,F6.2,1X,F6.2,1X,F6.2,1X,F6.2,1X,F6.2)
230 FORMAT (I6,1X,E10.4,1X,E10.4,1X,E10.4,1X,F6.2,1X,F6.2)
240 FORMAT (E10.4,1X,E10.4,1X,E10.4,1X,E10.4,1X,E10.4,1X,E10.4)
250 FORMAT (7(A8,1X))
260 FORMAT (F8.0,1X,4(F8.3,1X),2(F8.1,1X))

!End Main Program


1000 END

!*************************************************************************
! SUBROUTINES
!*************************************************************************
!------------------------------------------------------------------
!SUBROUTINE TO CALCULATE THE CONNECTIVITY VECTORS FOR EACH ELEMENT
!------------------------------------------------------------------
SUBROUTINE CONNVEC(EL, NPE, NEX, NEZ, ELNODES)

INTEGER NPE
INTEGER EL, NEX, NEZ, ELNODES(NPE)
INTEGER I, I1, I2

!Loop over number of elements in x direction


DO I = 1,NEX

!Calculate lower and upper z-direction elements for each element


!in the x direction
I1 = I + (I-1) * (NEZ-1)
I2 = I1 + NEZ-1

!Determine connectivity vector if the element lies within the


!upper and lower bounds
IF(EL >= I1 .AND. EL <= I2) THEN
ELNODES(1) = EL + I-1
ELNODES(2) = ELNODES(1) + NEZ + 1

185
ELNODES(3) = ELNODES(2) + 1
ELNODES(4) = ELNODES(1) + 1
END IF

END DO

END

!------------------------------------------------------------------
!SUBROUTINE TO CALCULATE NODAL X AND Z COORDINATES
!------------------------------------------------------------------
SUBROUTINE NODCOORD(RJ, THP, NEX, NEZ, NNX, NNZ, OFFSET,
* PADANGLE, L, ELNODES, XJ, ZJ)

INTEGER NEX, NEZ, NNX, NNZ


REAL*8 XJ, ZJ, RJ, ZS, ZE
REAL*8 L, OFFSET, PADANGLE
REAL*8 THS, THE, THP
INTEGER I, II, II1, II2, JJ, ELNODES
PARAMETER (PI = 3.14159265359)

THS = THP - OFFSET*PADANGLE ! Start angle of pads(radians)


THE = THP + (1 - OFFSET) * PADANGLE ! End angle of pads(radians)

ZS = -L/2 ! Starting Z-coordinate


ZE = L/2 ! Ending Z-coordinate

!Get global node number from the connectivity vector


I = ELNODES

DO II = 1,NNX

!Calculate lower and upper z-direction nodes for each node


!in the theta direction
II1 = (II-1) * NNZ + 1
II2 = II1 + (NNZ-1)

!Calculate the theta coordinate for the node


IF(I >= II1 .AND. I <= II2) THEN
XJ = RJ*( THS + (II-1)*PADANGLE/NEX )
END IF

END DO

DO JJ = 1,NNZ
!Calculate the z coordinate for the node
IF( MOD(I-JJ,NNZ) == 0 ) THEN
ZJ = ZS + (JJ-1)*L/NEZ
END IF
END DO

END

!------------------------------------------------------------------
!SUBROUTINE FOR THE MAXIMUM TILT ANGLE (DMAX)
!------------------------------------------------------------------
SUBROUTINE MAXDELTA(PADANGLE,M,OFFSET,THP,TP,RJ, ECC,CB,THC,DMAX)

186
REAL*8 M, CB, CP, RJ, RB, RP, ECC
REAL*8 OFFSET, PADANGLE, THC, TP
REAL*8 THP, THE, DMAX

CP = CB/(1-M) ! Radial pad clearance (in)


RB = RJ + CB ! Radius of "bushing" (in)
RP = RJ + CP ! Radius of pad curvature (in)
THE = THP + (1 - OFFSET) * PADANGLE ! End angle of pads(radians)

DMAX = ( (RP-RJ) + ECC*DCOS(THC-THE) - (RP-RB)*DCOS(THE-THP) )/


* ( (RP+TP)*DSIN(THE-THP) )

END

!------------------------------------------------------------------
!SUBROUTINE TO CALCULATE FILM THICKNESS EXPRESSIONS FOR A
!TILTING PAD JOURNAL BEARING WHICH HAS A CYLINDRICAL PIVOT.
!------------------------------------------------------------------
SUBROUTINE FILMTH(DELTA, M, THP,TP,RJ, ECC,CB,THC, XJ, DHTHJ, HJ)

REAL*8 M, CB, CP, RJ, RB, RP, ECC


REAL*8 THC, TP
REAL*8 THP, DELTA
REAL*8 TH, XJ, HJ, DHTHJ

!Do preliminary calculations using input parameters


CP = CB/(1-M) ! Radial pad clearance (in)
RB = RJ + CB ! Radius of "bushing" (in)
RP = RJ + CP ! Radius of pad curvature (in)

TH = XJ/RJ

!Calculate dh/dtheta for element nodes


DHTHJ = ECC*DSIN(THC-TH) + (RP-RB)*DSIN(TH-THP) -
* DELTA*(RP+TP)*DCOS(TH-THP)

!Calculate film thickness, h, for element nodes


HJ = (RP-RJ) + ECC*DCOS(THC-TH) - (RP-RB)*DCOS(TH-THP) -
* DELTA*(RP+TP)*DSIN(TH-THP)

END

!------------------------------------------------------------------
!SUBROUTINE TO CALCULATE BASIS FUNCTIONS PHI, DPHI/DX, DPHI/DZ
!AND THE DETERMINANT OF THE JACOBIAN
!------------------------------------------------------------------
SUBROUTINE BASFUNC(XI, ETA, XJ, ZJ, PH, DPHX, DPHZ, DETJA)

REAL*8 XI, ETA, XJ(4), ZJ(4)


REAL*8 PH(4), DPHXI(4), DPHETA(4)
REAL*8 DPHX(4), DPHZ(4), JA(2,2), DETJA
REAL*8 XIC(4), ETAC(4)
INTEGER I

!Define the xi and eta coefficients


XIC = (/-1, 1, 1, -1/)*1.0D0
ETAC = (/-1, -1, 1, 1/)*1.0D0

!Calculate the basis functions phi, dphi/dxi, dphi/deta


DO I = 1,4

187
PH(I) = (1 + XIC(I)*XI) * (1 + ETAC(I)*ETA)/4.0
DPHXI(I) = XIC(I) * (1 + ETAC(I)*ETA)/4.0
DPHETA(I) = ETAC(I) * (1 + XIC(I)*XI)/4.0
END DO

!Initialize the Jacobian


DO I = 1,2
DO J = 1,2
JA(I,J) = 0.0D0
END DO
END DO

!Calculate the components of the Jacobian


DO I = 1,4
JA(1,1) = JA(1,1) + XJ(I)*DPHXI(I)
JA(1,2) = JA(1,2) + XJ(I)*DPHETA(I)
JA(2,1) = JA(2,1) + ZJ(I)*DPHXI(I)
JA(2,2) = JA(2,2) + ZJ(I)*DPHETA(I)
END DO

!Calculate the determinant of the Jacobian


DETJA = JA(1,1)*JA(2,2) - JA(2,1)*JA(1,2)

!Calculate dphi/dx, dphi/dz


DO I = 1,4
DPHX(I) = ( JA(2,2)*DPHXI(I) - JA(2,1)*DPHETA(I) )/DETJA
DPHZ(I) = ( -JA(1,2)*DPHXI(I) + JA(1,1)*DPHETA(I) )/DETJA
END DO

END

!------------------------------------------------------------------
!SUBROUTINE TO PERFORM NUMERICAL INTEGRATION AND EVALUATE K AND
!F OVER A GIVEN ELEMENT
!------------------------------------------------------------------
SUBROUTINE ELMAT(VX, VY, VD, TP, THP, NPE, EDOF, XJ, ZJ, HJ,
* DHTHJ, MU, RJ, RP, U, KE, FE)

INTEGER NPE, EDOF, NG


PARAMETER (NG = 4)
REAL*8 XJ(NPE), ZJ(NPE), HJ(NPE), DHTHJ(NPE)
REAL*8 MU, RJ, RP, U, VX, VY, VD, THP, TP
REAL*8 XI(NPE), ETA(NPE), W(NPE), H, DHTH, TH
REAL*8 FE(EDOF), KE(EDOF,EDOF)
REAL*8 A1, A2
REAL*8 PH(EDOF), DPHX(EDOF), DPHZ(EDOF), DETJA
INTEGER I, J, M

!Define Gauss points and weights


XI = (/-1.0D0, 1.0D0, 1.0D0, -1.0D0/) / DSQRT(3.0D0)
ETA = (/-1.0D0, -1.0D0, 1.0D0, 1.0D0/) / DSQRT(3.0D0)
W = (/1.0D0, 1.0D0, 1.0D0, 1.0D0/)

!Initialize FE and KE.


DO I = 1,EDOF
FE(I) = 0.0D0

DO J = 1,EDOF
KE(I,J) = 0.0D0

188
END DO
END DO

!Begin Gauss point loop


DO M = 1,NG

!Get basis function terms


CALL BASFUNC(XI(M), ETA(M), XJ, ZJ, PH, DPHX, DPHZ, DETJA)

!Assemble for pressure calculations

!Calculate h,th,dh/dtheta for the element


H = 0.0D0
TH = 0.0D0
DHTH = 0.0D0

DO I = 1,EDOF
H = H + PH(I)*HJ(I)
TH = TH + PH(I)*XJ(I)/RJ
DHTH = DHTH + PH(I)*DHTHJ(I)
END DO

!Calculate coefficients
A1 = H**(3.0D0) / MU
A2 = 6.0D0*U*DHTH/RJ - 12.0D0*(VX*DCOS(TH) + VY*DSIN(TH) +
* VD*(RP+TP)*DSIN(TH-THP))

!Loop over i, j and assemble FE and KE


DO I = 1,EDOF

FE(I) = FE(I) + W(M)*DETJA*( -A2 * PH(I) )

DO J = 1,EDOF

KE(I,J) = KE(I,J) + W(M)*DETJA*


* A1*( DPHX(I)*DPHX(J) + DPHZ(I)*DPHZ(J) )

END DO
END DO

!End Gauss point loop


END DO

END

!------------------------------------------------------------------
!SUBROUTINE TO ASSEMBLE THE BANDED GLOBAL STIFFNESS MATRIX (KG)
!AND THE GLOBAL LOAD VECTOR FG. CODE ADAPTED FROM J.N. REDDY,
!"AN INTRODUCTION TO THE FINITE ELEMENT METHOD", (1993), P.649
!------------------------------------------------------------------
SUBROUTINE GLOBMATBW(HBW, NPE, EDOF, NDOF, GDOF, ELNODES,
* FE, KE, FG, KG)

INTEGER EDOF, NDOF, GDOF, NPE, HBW


REAL*8 FE(EDOF), KE(EDOF,EDOF), FG(GDOF), KG(GDOF,HBW)
INTEGER ELNODES(NPE)
INTEGER I, J, II, JJ, NR, NC, NCL, M

189
DO I = 1,NPE
NR = (ELNODES(I)-1)*NDOF

DO II = 1,NDOF
NR = NR + 1
L = (I-1)*NDOF + II

FG(NR) = FG(NR) + FE(L)

DO J = 1,NPE
NCL = (ELNODES(J)-1)*NDOF

DO JJ = 1,NDOF
M = (J-1)*NDOF + JJ
NC = NCL + JJ + 1 - NR

IF (NC > 0) THEN


KG(NR,NC) = KG(NR,NC) + KE(L,M)
END IF
END DO

END DO

END DO

END DO

END

!------------------------------------------------------------------
!SUBROUTINE TO GENERATE THE BOUNDARY NODES
!------------------------------------------------------------------
SUBROUTINE BNODES(NNX, NNZ, NBN, BOUNOD)

INTEGER NNX, NNZ, NBN


INTEGER S1NOD(NNZ), S2NOD(NNX-2), S3NOD(NNZ), S4NOD(NNX-2)
INTEGER BOUNOD(NBN)
INTEGER I

DO I = 1,NNZ
S1NOD(I) = I
S3NOD(I) = NNZ*(NNX-1) + I
BOUNOD(I) = S1NOD(I)
BOUNOD(NNZ + I) = S3NOD(I)
END DO

DO I = 1,NNX-2
S2NOD(I) = 2*NNZ + (I-1)*NNZ
S4NOD(I) = (NNZ+1) + (I-1)*NNZ
BOUNOD(2*NNZ + I) = S2NOD(I)
BOUNOD(2*NNZ + (NNX-2) + I) = S4NOD(I)
END DO

END

!------------------------------------------------------------------
!SUBROUTINE TO APPLY BOUNDARY CONDITIONS. CODE ADAPTED FROM J.N.
!REDDY,"AN INTRODUCTION TO THE FINITE ELEMENT METHOD",(1993),P.654
!------------------------------------------------------------------

190
SUBROUTINE APPLYBCSBW(FG, KG, HBW, NDOF, GDOF, NBN, NNX, NNZ,VAL1)

INTEGER NDOF, GDOF, NBN, HBW, NNX, NNZ, NPE,EDOF


PARAMETER (NPE = 4,EDOF = 4)
REAL*8 FG(GDOF), KG(GDOF,HBW)
REAL*8 VALUE, VAL1
INTEGER BOUDOF(NBN), BOUNOD(NBN)
INTEGER I, II, J, IE, IT, NB

!Apply essential BCs


DO I = 1,NBN
BOUDOF(I) = 1
END DO

CALL BNODES(NNX, NNZ, NBN, BOUNOD)

DO NB = 1,NBN
IE = (BOUNOD(NB) - 1)*NDOF + BOUDOF(NB)
VALUE = VAL1

IT = HBW - 1
I = IE - HBW

DO II = 1,IT
I = I + 1

IF (I >= 1 ) THEN
J = IE - I + 1
FG(I) = FG(I) - KG(I,J)*VALUE
KG(I,J) = 0.0
END IF
END DO

KG(IE,1) = 1.0
FG(IE) = VALUE
I = IE

DO II = 2,HBW
I = I + 1

IF (I <= GDOF) THEN


FG(I) = FG(I) - KG(IE,II)*VALUE
KG(IE,II) = 0.0
END IF
END DO

END DO

END
C ________________________________________________________________
C
C The subroutine solves banded symmetric system of equations using
C gauss elimination. The banded matrix is input as BAND(NEQNS,NBW)
C and the right-hand side is input as RHS(NEQNS), where NBW is the
C half band width and NEQNS is the number of equations. If IRES is
C greater than zero, the right hand elimination is skipped.
C CODE FROM J.N. REDDY,"AN INTRODUCTION TO THE FINITE ELEMENT
C METHOD",(1993),P.676
C ________________________________________________________________

191
SUBROUTINE SOLVEBW(NRM,NCM,NEQNS,NBW,BAND,RHS,IRES)

IMPLICIT REAL*8(A-H,O-Z)
DIMENSION BAND(NRM,NCM),RHS(NRM)
COMMON/IO/IN,ITT
MEQNS=NEQNS-1
IF(IRES.LE.0) THEN
DO 30 NPIV=1,MEQNS
NPIVOT=NPIV+1
LSTSUB=NPIV+NBW-1
IF(LSTSUB.GT.NEQNS) THEN
LSTSUB=NEQNS
ENDIF
DO 20 NROW=NPIVOT,LSTSUB
C
C INVERT ROWS AND COLUMNS FOR ROW FACTOR
C
NCOL=NROW-NPIV+1
FACTOR=BAND(NPIV,NCOL)/BAND(NPIV,1)
DO 10 NCOL=NROW,LSTSUB
ICOL=NCOL-NROW+1
JCOL=NCOL-NPIV+1
10 BAND(NROW,ICOL)=BAND(NROW,ICOL)-FACTOR*BAND(NPIV,JCOL)
20 RHS(NROW)=RHS(NROW)-FACTOR*RHS(NPIV)
30 CONTINUE
ELSE
40 DO 60 NPIV=1,MEQNS
NPIVOT=NPIV+1
LSTSUB=NPIV+NBW-1
IF(LSTSUB.GT.NEQNS) THEN
LSTSUB=NEQNS
ENDIF
DO 50 NROW=NPIVOT,LSTSUB
NCOL=NROW-NPIV+1
FACTOR=BAND(NPIV,NCOL)/BAND(NPIV,1)
50 RHS(NROW)=RHS(NROW)-FACTOR*RHS(NPIV)
60 CONTINUE
ENDIF
C
C Back substitution
C
DO 90 IJK=2,NEQNS
NPIV=NEQNS-IJK+2
RHS(NPIV)=RHS(NPIV)/BAND(NPIV,1)
LSTSUB=NPIV-NBW+1
IF(LSTSUB.LT.1) THEN
LSTSUB=1
ENDIF
NPIVOT=NPIV-1
DO 80 JKI=LSTSUB,NPIVOT
NROW=NPIVOT-JKI+LSTSUB
NCOL=NPIV-NROW+1
FACTOR=BAND(NROW,NCOL)
80 RHS(NROW)=RHS(NROW)-FACTOR*RHS(NPIV)
90 CONTINUE
RHS(1)=RHS(1)/BAND(1,1)
RETURN
END

!------------------------------------------------------------------

192
!SUBROUTINE TO SOLVE THE NON-BANADED LINEAR SYSTEM DX = B USING
!GAUSSIAN ELIMINATION WITH PARTIAL PIVOTING - ALGORITHM 6.2 PAGE 340,
!R.L. BURDEN, J.D. FAIRES, NUMERICAL ANALYSIS, FIFTH EDITION, 1993
!------------------------------------------------------------------
SUBROUTINE SOLVE(D, B, N, X)

REAL*8 D(N,N), B(N), X(N), A(N,N+1), AP, M, SUM, ATEMP


INTEGER I, J, P, K

!Load D into A
DO I = 1,N
DO J = 1,N
A(I,J) = D(I,J)
END DO
END DO

!Load B into N+1 column of A


DO I = 1,N
A(I,N+1) = B(I)
END DO

!Begin Gaussian elimination


DO I = 1,N-1
P = I
AP = A(I,I)

!Find the maximum pivot value in the Ith column and the
!corresponding row number
DO J = I+1,N
IF ( DABS(A(J,I)) > DABS(AP) ) THEN
AP = A(J,I)
P = J
END IF
END DO

!Quit if the pivot value is zero


IF (AP == 0) THEN
PRINT *, 'NO UNIQUE SOLUTION - PIVOT = 0'
PRINT *, I, P, AP
GO TO 1000
END IF

!Switch I and Pivot rows in A matrix


DO K = 1,N+1
ATEMP = A(I,K)
A(I,K) = A(P,K)
A(P,K) = ATEMP
END DO

!Do pivoting with maximum value in Ith column


DO J = I+1,N
M = A(J,I)/(1.0D0*AP)

DO K = 1,N+1
A(J,K) = A(J,K) - M*A(I,K)
END DO
END DO
END DO

193
!Quit if denominator for calculating xn is zero
IF ( A(N,N) == 0 ) THEN
PRINT *, 'NO UNIQUE SOLUTION - XN DENOMINATOR = 0'
PRINT *, N
PRINT *, (A(I,N), I = 1,N)
GO TO 1000
END IF

!Back substitution
X(N) = A(N,N+1)/(1.0D0*A(N,N))

DO K = 1,N-1

I = N-K
SUM = 0

DO J = I+1,N
SUM = SUM + A(I,J)*X(J)
END DO

X(I) = ( A(I,N+1) - SUM ) /(1.0D0*A(I,I))


END DO

1000 END

!------------------------------------------------------------------
!SUBROUTINE TO CALCULATE MOMENT
!------------------------------------------------------------------
SUBROUTINE MOCALC(GDOF, EDOF, NEX, NEZ, NE, NNX, NNZ,
* NPE, L, PADANGLE, OFFSET, THP, TP,
* RJ, PRESS, MOMENT)

INTEGER GDOF, EDOF, NEX, NEZ, NE, NNX, NNZ, NPE, ELNODES(NPE),NG
PARAMETER (NG = 4)
REAL*8 XJ(EDOF), ZJ(EDOF), PJ(EDOF), MOMENT
REAL*8 XI(NG), ETA(NG), W(NG), P, PRESS(GDOF)
REAL*8 L, PADANGLE, OFFSET, THP, TP, RJ
REAL*8 PH(EDOF), DPHX(EDOF), DPHZ(EDOF), DETJA, TH
INTEGER I, J, EL

!Define Gauss points and weights


XI = (/-1.0D0, 1.0D0, 1.0D0, -1.0D0/) / DSQRT(3.0D0)
ETA = (/-1.0D0, -1.0D0, 1.0D0, 1.0D0/) / DSQRT(3.0D0)
W = (/1.0D0, 1.0D0, 1.0D0, 1.0D0/)

!Initialize moment
MOMENT = 0.0D0

!Begin element loop


DO EL = 1, NE

!Calculate the element connectivity vector


CALL CONNVEC(EL, NPE, NEX, NEZ, ELNODES)

!Get the global coordinates for the element nodes


DO I = 1,NPE
CALL NODCOORD(RJ, THP, NEX, NEZ, NNX, NNZ, OFFSET,
* PADANGLE, L, ELNODES(I), XJ(I), ZJ(I))
END DO

194
!Calculate the nodal moment arms and pressures
DO I = 1,NPE
PJ(I) = PRESS(ELNODES(I))
END DO

!Begin Gauss point loop


DO J = 1,NG

!Get basis function terms


CALL BASFUNC(XI(J), ETA(J), XJ, ZJ,
* PH, DPHX, DPHZ, DETJA)

!Calculate the moment arm and pressure for the element


P = 0
TH = 0

DO I = 1,EDOF
P = P + PH(I)*PJ(I)
TH = TH + PH(I)*XJ(I)/RJ
END DO

!Assemble MOMENT
MOMENT = MOMENT + W(J)*DETJA*(RJ+TP)*P*
* (TH-THP)*DCOS(TH-THP)

!End Gauss point loop


END DO

!End element loop


END DO

END

!------------------------------------------------------------------
!SUBROUTINE TO CALCULATE THE STATIC LOAD FOR A PAD
!------------------------------------------------------------------
SUBROUTINE PLOADCALC(NE, NPE, EDOF, GDOF, NEX, NEZ, NNX, NNZ,
* RJ, THP, OFFSET, PADANGLE, L, PRESS,
* PLOADX, PLOADY)

INTEGER NE, EDOF, GDOF, NPE, NEX, NEZ, NNX, NNZ, ELNODES(NPE)
REAL*8 XJ(EDOF), ZJ(EDOF), PJ(EDOF), PLOADX, PLOADY, RJ
REAL*8 XI(EDOF), ETA(EDOF), W(EDOF), P, TH, THP
REAL*8 OFFSET, PADANGLE, L, PRESS(GDOF)
REAL*8 PH(EDOF), DPHX(EDOF), DPHZ(EDOF), DETJA
INTEGER I, M, EL

!Define Gauss points and weights


XI = (/-1.0D0, 1.0D0, 1.0D0, -1.0D0/) / DSQRT(3.0D0)
ETA = (/-1.0D0, -1.0D0, 1.0D0, 1.0D0/) / DSQRT(3.0D0)
W = (/1.0D0, 1.0D0, 1.0D0, 1.0D0/)

!Initialize static load components


PLOADX = 0.0D0
PLOADY = 0.0D0

DO EL = 1,NE
!CALCULATE THE CONNECTIVITY VECTOR FOR THE ELEMENT
CALL CONNVEC(EL, NPE, NEX, NEZ, ELNODES)

195
!GET THE GLOBAL COORDINATES FOR EACH NODE IN THE CONNECTIVITY
VECTOR
DO I = 1,NPE
CALL NODCOORD(RJ, THP, NEX, NEZ, NNX, NNZ, OFFSET,
* PADANGLE, L, ELNODES(I), XJ(I), ZJ(I))
END DO

!CALCULATE THE PRESSURE AT EACH NODE


DO I = 1,NPE
PJ(I) = PRESS(ELNODES(I))
END DO

!INTEGRATE OVER PAD TO GET STATIC LOAD COMPONENTS


!Begin Gauss point loop
DO M = 1,EDOF

!Get basis function terms


CALL BASFUNC(XI(M), ETA(M), XJ, ZJ,
* PH, DPHX, DPHZ, DETJA)

!Calculate the pressure and theta for the element


P = 0.0D0
TH = 0.0D0

DO I = 1,EDOF
P = P + PH(I)*PJ(I)
TH = TH + PH(I)*XJ(I)/RJ
END DO

!Calculate the static load components


PLOADX = PLOADX + W(M)*DETJA*P*DCOS(TH)
PLOADY = PLOADY + W(M)*DETJA*P*DSIN(TH)

END DO

END DO

END

!------------------------------------------------------------------
!SUBROUTINE TO PERTURB THE JOURNAL FOR THE CALCULATION OF THE
!STIFFNESS AND DAMPING COEFFICIENTS
!------------------------------------------------------------------
SUBROUTINE PERTURB(PERTFLAG, PI, DECC, DV, EQECC, EQTHC,
* ECC, THC, VX, VY)

REAL*8 PI, ECC, THC, VX, VY, DVX, DVY, EQECC, EQTHC
REAL*8 ECCX, ECCY, DECCX, DECCY, DECC, DV
INTEGER PERTFLAG

SELECT CASE (PERTFLAG)

!Perturb eccx -> Kxx, Kyx


CASE(1)
DECCX = DECC
ECCX = DECCX - EQECC*DCOS(EQTHC)
ECCY = -EQECC*DSIN(EQTHC)
ECC = DSQRT(ECCX**2 + ECCY**2)
THC = PI + DATAN2(ECCY,ECCX)

196
!Perturb eccy -> Kyy, Kxy
CASE(2)
DECCY = DECC
ECCX = - EQECC*DCOS(EQTHC)
ECCY = DECCY - EQECC*DSIN(EQTHC)
ECC = DSQRT(ECCX**2 + ECCY**2)
THC = PI + DATAN2(ECCY,ECCX)

!Perturb Vx -> Cxx, Cyx


CASE(3)
DVX = DV
VX = VX + DVX
ECC = EQECC
THC = EQTHC

!Perturb Vy -> Cyy, Cxy


CASE(4)
DVY = DV
VY = VY + DVY
ECC = EQECC
THC = EQTHC

END SELECT

END

!------------------------------------------------------------------
!SUBROUTINE TO CALCULATE STIFFNESS AND DAMPING COEFFICIENTS
!------------------------------------------------------------------
SUBROUTINE SDCOEFF(MAXITER,GDOF, EDOF, NBN, NEX, NEZ, NE, NNX,
* NNZ, NDOF, HBW, NPE, NPADS, U, MU, P0, L, PADANGLE, M,OFFSET,
* THETAP, TP, TOL2, DXX, VX, LOAD, VY, VD, RJ, RP, CB, OMEGA,
* EQECC, EQTHC, EQSFX, EQSFY, RATIO, PI, KIJ, CIJ)

INTEGER GDOF, EDOF, NBN, NEX, NEZ, NE, NNX, NNZ, NDOF,
* HBW, NPE, NPADS, PERTFLAG, MAXITER
REAL*8 L, PADANGLE, M, OFFSET, THETAP(NPADS),
* TP, TOL2, DXX, LOAD, FX(NPADS), FY(NPADS), DEL(NPADS),
* RJ, EQECC, CB, PI, EQTHC, RP, U, MU, P0, MOMENT(NPADS),
* VX, VY, VD, EQSFX, EQSFY, SFX, SFY, OMEGA, RATIO,
* ECC, THC, DECC, DV, KIJ(4), CIJ(4)

DECC = RATIO*CB
DV = OMEGA*DECC

DO PERTFLAG = 1,4
VX = 0.0D0
VY = 0.0D0
VD = 0.0D0

CALL PERTURB(PERTFLAG, PI, DECC, DV, EQECC, EQTHC,


* ECC, THC, VX, VY)

CALL NEWTON(GDOF, EDOF, NBN, NEX, NEZ, NE, NNX, NNZ, NPADS,
* NDOF, HBW, NPE, NPADS,MAXITER,PERTFLAG,TOL2,DXX,
* L,PADANGLE,M, OFFSET, THETAP, TP, RJ, RP, CB,
* P0, VX, VY, VD, LOAD,
* U, MU, SFX, SFY,FX, FY, DEL, MOMENT,ECC, THC)

SELECT CASE (PERTFLAG)

197
CASE(1)
KIJ(1) = (SFX-EQSFX)/DECC !Kxx
KIJ(2) = (SFY-EQSFY)/DECC !Kyx

CASE(2)
KIJ(3) = (SFX-EQSFX)/DECC !Kxy
KIJ(4) = (SFY-EQSFY)/DECC !Kyy

CASE(3)
CIJ(1) = (SFX-EQSFX)/DV !Cxx
CIJ(2) = (SFY-EQSFY)/DV !Cyx

CASE(4)
CIJ(3) = (SFX-EQSFX)/DV !Cxy
CIJ(4) = (SFY-EQSFY)/DV !Cyy

END SELECT

END DO

END

!------------------------------------------------------------------
!SUBROUTINE TO CALCULATE THE PAD PRESSURE DISTRIBUTION
!------------------------------------------------------------------
SUBROUTINE PRESSCALC(GDOF, EDOF, NBN, NEX, NEZ, NE, NNX, NNZ,
* NDOF, HBW, NPE, U, MU, P0, L, PADANGLE, M,OFFSET,
* THP, TP, VX, VY, VD, RJ, RP,ECC, CB, DELTA, THC, PRESS)

INTEGER GDOF, EDOF, NBN, NEX, NEZ, NE, NNX, NNZ, NDOF,
* HBW, NPE, ELNODES(NPE), I, J, EL
REAL*8 KE(EDOF,EDOF), FE(EDOF), KG(GDOF,HBW), FG(GDOF),PRESS(GDOF)
REAL*8 L, PADANGLE, M, OFFSET, THP, TP, RJ, ECC, CB, THC, RP
REAL*8 U, MU, P0, DELTA
REAL*8 XJ(NPE), ZJ(NPE), DHTHJ(NPE), HJ(NPE), VX, VY, VD

!Initialize Global Matrices


DO I = 1,GDOF
FG(I) = 0

DO J = 1,HBW
KG(I,J) = 0
END DO
END DO

!Begin element loop


DO EL = 1, NE

!Calculate the element connectivity vector


CALL CONNVEC(EL, NPE, NEX, NEZ, ELNODES)

!Get the global coordinates for the element nodes


DO I = 1,NPE
CALL NODCOORD(RJ, THP, NEX, NEZ, NNX, NNZ, OFFSET,
* PADANGLE, L, ELNODES(I), XJ(I), ZJ(I))
END DO

!Calculate the film thickness


DO I = 1,NPE

198
CALL FILMTH(DELTA, M,THP,TP,RJ,
* ECC,CB,THC, XJ(I), DHTHJ(I), HJ(I))
END DO

!Calculate the element stiffness matrix(KE) and load vector(FE)


CALL ELMAT(VX, VY, VD, TP, THP, NPE, EDOF, XJ, ZJ, HJ,
* DHTHJ, MU, RJ, RP, U, KE, FE)

!Assemble the stiffness matrix(KG) and load vector(FG)


CALL GLOBMATBW(HBW, NPE, EDOF, NDOF, GDOF, ELNODES,
* FE, KE, FG, KG)

!End element loop


END DO

!Apply essential boundary conditions


CALL APPLYBCSBW(FG, KG, HBW, NDOF, GDOF, NBN, NNX, NNZ, P0)

!Solve system of equations


CALL SOLVEBW(GDOF,HBW,GDOF,HBW,KG,FG,0)

!Place the calculated pressure values into vector PRESS


DO I = 1,GDOF
PRESS(I) = FG(I)

!Cavitation boundary condition


IF(PRESS(I) < P0) THEN
PRESS(I) = P0
END IF

END DO

END

!------------------------------------------------------------------
!SUBROUTINE TO CALCULATE THE PAD TEMPERATURE DISTRIBUTION
!------------------------------------------------------------------
SUBROUTINE TEMPCALC(GDOF, NEX, NEZ, NNX, NNZ, U, MU, RHOL, CAPL,
* T0, HTC, TIN, TAMB, L, PADANGLE, M,OFFSET,
* THP, TP, RJ, ECC, CB, DELTA, THC, TEMP)

INTEGER GDOF, NEX, NEZ, NNX, NNZ,


* I, J, I1, I2
REAL*8 L, PADANGLE, M, OFFSET, THP, TP, RJ, ECC, CB, THC
REAL*8 U, MU, RHOL, CAPL, T0, HTC, TIN, TAMB
REAL*8 DELTA,TEMP(GDOF), TC(NNX),X, Z, DTH,TH, THS,A,B
REAL*8 DHTH, H, OMEGA

!Calculate element dimensions


DTH = PADANGLE/(1.0D0*NEX)
THS = THP - OFFSET*PADANGLE ! Start angle of pads(radians)
OMEGA = U/RJ

!Initialize
DO I = 1,NNX
TC(I) = TIN
END DO

199
DO I = 1,NNX-1

TH = THS + 1.0D0*I*DTH
X = RJ*TH

CALL FILMTH(DELTA, M,THP,TP,RJ,ECC,CB,THC, X, DHTH, H)

A = 2.0D0*HTC/(RHOL*CAPL*OMEGA*H)
B = A*TAMB + 2.0D0*MU*OMEGA*RJ**2.0/(RHOL*CAPL*H**2.0)

TC(I+1) = (1.0D0-DTH*A)*TC(I) + DTH*B


END DO

DO I = 1, NNX

!Calculate lower and upper z-direction nodes for each node


!in the theta direction
I1 = (I-1) * NNZ + 1
I2 = I1 + (NNZ-1)

DO J = I1,I2

!Get the z-coordinate for the global node


CALL NODCOORD(RJ, THP, NEX, NEZ, NNX, NNZ, OFFSET,
* PADANGLE, L, J, X, Z)

!Calculate the temperature profile assuming a parabolic


!axial distribution
TEMP(J) = 4*(T0 - TC(I))*Z**2/L**2 + TC(I)
END DO

END DO

END

!------------------------------------------------------------------
!SUBROUTINE TO CALCULATE XX - VECTOR OF TILT ANGLES AND JOURNAL POS.
!------------------------------------------------------------------
SUBROUTINE XXCALC(NPADS,PADANGLE,M,OFFSET,THETAP,TP,RJ,ECC,
* PERTFLAG, CB,THC, XXDOF, XX)

INTEGER NPADS, XXDOF, PERTFLAG, I


REAL*8 PADANGLE,M,OFFSET,THP,THETAP(NPADS), TP,RJ,ECC,CB,THC
REAL*8 XX(XXDOF),DMAX(NPADS)

DO I = 1,NPADS
THP = THETAP(I)

CALL MAXDELTA(PADANGLE,M,OFFSET,THP,TP,RJ,ECC,CB,THC,DMAX(I))

XX(I) = 0.5*DMAX(I)
END DO

IF (PERTFLAG == 0) THEN
XX(XXDOF-1) = ECC
XX(XXDOF) = THC
END IF

200
END

!------------------------------------------------------------------
!SUBROUTINE TO USE NEWTON'S METHOD TO SOLVE A SYSTEM OF EQUATIONS
!------------------------------------------------------------------
SUBROUTINE NEWTON(GDOF, EDOF, NBN, NEX, NEZ, NE, NNX, NNZ, NPADS,
* NDOF, HBW, NPE, XXDOF,MAXITER,PERTFLAG,TOL,
DXX,L,
* PADANGLE,M, OFFSET, THETAP, TP, RJ, RP, CB,
* P0, VX, VY, VD, LOAD,
* U, MU, SFX, SFY,FX, FY, DEL, MOMENT,ECC, THC)

INTEGER GDOF, EDOF, NBN, NEX, NEZ, NE, NNX, NNZ, NPADS, NDOF,
* HBW, NPE, XXDOF, MAXITER
REAL*8 TOL, DXX, PRESS(GDOF)
REAL*8 L, PADANGLE, M, OFFSET, THP, THETAP(NPADS), TP
REAL*8 RJ, RP, ECC, CB, THC, LOAD
REAL*8 U, MU, P0, DELTA, DEL(NPADS)
REAL*8 VX, VY, VD
REAL*8 FX(NPADS), FY(NPADS), SFX,SFY,MOMENT(NPADS)
REAL*8 FXA(NPADS), FYA(NPADS), SFXA,SFYA,MOMENTA(NPADS)
REAL*8 XX(XXDOF), FF(XXDOF), XXA(XXDOF), FFA(XXDOF)
REAL*8 ECCA, THCA
REAL*8 JAC(XXDOF,XXDOF), YY(XXDOF), YYMAX
INTEGER NITER, PADNUM, CONFLAG, PERTFLAG
INTEGER I, J, K

CONFLAG = 0 !Initialize convergence flag


NITER = 1 !Set counter to first iteration

!DO ONLY FOR EQUILIBRIUM CALCULATION


IF (PERTFLAG == 0) THEN

!Initialize XX
CALL XXCALC(NPADS,PADANGLE,M,OFFSET,THETAP,TP,RJ,ECC,
* PERTFLAG, CB,THC, XXDOF, XX)

!Indicate perturbation type


PRINT *, 'PERTFLAG = ',PERTFLAG

!Begin Newton's method iteration loop


DO WHILE (CONFLAG .EQ. 0 .AND. NITER .LE. MAXITER)

DO PADNUM = 1,NPADS

!Get preliminary inputs for each pad


THP = THETAP(PADNUM)
DELTA = XX(PADNUM)
ECC = XX(XXDOF-1)
THC = XX(XXDOF)

!Calculate pad pressure distributions


CALL PRESSCALC(GDOF, EDOF, NBN, NEX, NEZ, NE, NNX, NNZ,
* NDOF, HBW, NPE, U, MU, P0, L, PADANGLE, M,OFFSET,
* THP, TP, VX, VY, VD, RJ, RP,ECC, CB, DELTA, THC, PRESS)

!Calculate pad moments

201
CALL MOCALC(GDOF, EDOF, NEX, NEZ, NE, NNX, NNZ,
* NPE, L, PADANGLE, OFFSET, THP, TP,
* RJ, PRESS, MOMENT(PADNUM))

!Calculate the x,y force components ON each pad


CALL PLOADCALC(NE, NPE, EDOF, GDOF, NEX, NEZ, NNX, NNZ,
* RJ, THP, OFFSET, PADANGLE, L, PRESS,
* FX(PADNUM), FY(PADNUM))

!Store tilt angles


DEL(PADNUM) = DELTA

END DO

!Add up x and y forces on the pads


SFX = SUM(FX)
SFY = SUM(FY)

!Calculate FF
DO PADNUM = 1,NPADS
FF(PADNUM) = MOMENT(PADNUM)
END DO

FF(XXDOF-1) = SFX
FF(XXDOF) = DABS(SFY)-LOAD

!Print out iteration number, FF


! PRINT 10, NITER, (FF(I), I = 1,XXDOF)
PRINT *, 'iteration # = ', NITER

!Increment XX and store as XXA


DO I = 1,XXDOF
DO J = 1,XXDOF

DO K = 1,XXDOF
XXA(K) = XX(K)
END DO

XXA(J) = XX(J) + DXX

DO PADNUM = 1,NPADS

!Get preliminary inputs for each pad


THP = THETAP(PADNUM)
DELTA = XXA(PADNUM)
ECCA = XXA(XXDOF-1)
THCA = XXA(XXDOF)

!Calculate pad pressure distributions


CALL PRESSCALC(GDOF, EDOF, NBN, NEX, NEZ, NE, NNX, NNZ,
* NDOF, HBW, NPE, U, MU, P0, L, PADANGLE, M,OFFSET,
* THP, TP, VX, VY, VD, RJ, RP,ECCA, CB, DELTA, THCA, PRESS)

!Calculate pad moments


CALL MOCALC(GDOF, EDOF, NEX, NEZ, NE, NNX, NNZ,
* NPE, L, PADANGLE, OFFSET, THP, TP,
* RJ, PRESS, MOMENTA(PADNUM))

202
!Calculate the x,y force components ON each pad
CALL PLOADCALC(NE, NPE, EDOF, GDOF, NEX, NEZ,
* NNX, NNZ,RJ, THP, OFFSET, PADANGLE, L, PRESS,
* FXA(PADNUM), FYA(PADNUM))

END DO

!Add up x and y forces on the pads


SFXA = SUM(FXA)
SFYA = SUM(FYA)

!Calculate FFA
DO PADNUM = 1,NPADS
FFA(PADNUM) = MOMENTA(PADNUM)
END DO

FFA(XXDOF-1) = SFXA
FFA(XXDOF) = DABS(SFYA)-LOAD

!Calculate the Jacobian


JAC(I,J) = (FFA(I)-FF(I))/DXX
END DO
END DO

!Solve for YY = -INV(JAC)*F


CALL SOLVE(JAC, -FF, XXDOF, YY)

!Update XX(K) = XX(K-1) - INV(JAC)*F = XX(K-1) + YY(K-1)


DO I = 1,XXDOF
XX(I) = XX(I) + YY(I)
END DO

!Calculate maximum difference between successive XX iterations


YYMAX = 0
DO I = 1,XXDOF
IF (DABS(YY(I)) > YYMAX) THEN
YYMAX = DABS(YY(I))
END IF
END DO

!Stop iteration and exit loop if there is a solution


IF (YYMAX < TOL) THEN
CONFLAG = 1
END IF

!Go to next iteration if no solution was found


NITER = NITER + 1

!End iteration loop


END DO

!Indicate whether MAXITER is too low


IF (NITER > MAXITER) THEN
PRINT *, 'MAXIMUM NUMBER OF ITERATIONS EXCEEDED'
END IF

!DO FOR PERTURBATIONS


ELSE

203
!Initialize XX
CALL XXCALC(NPADS,PADANGLE,M,OFFSET,THETAP,TP,RJ,ECC,
* PERTFLAG, CB,THC, XXDOF, XX)

!Indicate perturbation type


PRINT *, 'PERTFLAG = ',PERTFLAG

!Begin Newton's method iteration loop


DO WHILE (CONFLAG .EQ. 0 .AND. NITER .LE. MAXITER)

DO PADNUM = 1,NPADS

!Get preliminary inputs for each pad


THP = THETAP(PADNUM)
DELTA = XX(PADNUM)

!Calculate pad pressure distributions


CALL PRESSCALC(GDOF, EDOF, NBN, NEX, NEZ, NE, NNX, NNZ,
* NDOF, HBW, NPE, U, MU, P0, L, PADANGLE, M,OFFSET,
* THP, TP, VX, VY, VD, RJ, RP,ECC, CB, DELTA, THC, PRESS)

!Calculate pad moments


CALL MOCALC(GDOF, EDOF, NEX, NEZ, NE, NNX, NNZ,
* NPE, L, PADANGLE, OFFSET, THP, TP,
* RJ, PRESS, MOMENT(PADNUM))

!Store tilt angles


DEL(PADNUM) = DELTA

END DO

!Calculate FF
DO PADNUM = 1,NPADS
FF(PADNUM) = MOMENT(PADNUM)
END DO

PRINT *, 'iteration # = ', NITER


! PRINT 10, NITER, (FF(I), I = 1,XXDOF)

!Increment XX and store as XXA


DO I = 1,XXDOF
DO J = 1,XXDOF

DO K = 1,XXDOF
XXA(K) = XX(K)
END DO

XXA(J) = XX(J) + DXX

DO PADNUM = 1,NPADS

!Get preliminary inputs for each pad


THP = THETAP(PADNUM)
DELTA = XXA(PADNUM)

!Calculate pad pressure distributions


CALL PRESSCALC(GDOF, EDOF, NBN, NEX, NEZ, NE, NNX, NNZ,
* NDOF, HBW, NPE, U, MU, P0, L, PADANGLE, M,OFFSET,
* THP, TP, VX, VY, VD, RJ, RP,ECC, CB, DELTA, THC, PRESS)

204
!Calculate pad moments
CALL MOCALC(GDOF, EDOF, NEX, NEZ, NE, NNX, NNZ,
* NPE, L, PADANGLE, OFFSET, THP, TP,
* RJ, PRESS, MOMENTA(PADNUM))

END DO

!Calculate FFA
DO PADNUM = 1,NPADS
FFA(PADNUM) = MOMENTA(PADNUM)
END DO

!Calculate the Jacobian


JAC(I,J) = (FFA(I)-FF(I))/DXX
END DO
END DO

!Solve for YY = -INV(JAC)*F


CALL SOLVE(JAC, -FF, XXDOF, YY)

!Update XX(K) = XX(K-1) - INV(JAC)*F = XX(K-1) + YY(K-1)


DO I = 1,XXDOF
XX(I) = XX(I) + YY(I)
END DO

!Calculate maximum difference between successive XX iterations


YYMAX = 0
DO I = 1,XXDOF
IF (DABS(YY(I)) > YYMAX) THEN
YYMAX = DABS(YY(I))
END IF
END DO

!Stop iteration and exit loop if there is a solution


IF (YYMAX < TOL) THEN
CONFLAG = 1
END IF

!Go to next iteration if no solution was found


NITER = NITER + 1

!End iteration loop


END DO

!Indicate whether MAXITER is too low


IF (NITER > MAXITER) THEN
PRINT *, 'MAXIMUM NUMBER OF ITERATIONS EXCEEDED'

ELSE

DO PADNUM = 1,NPADS

!Get preliminary inputs for each pad


THP = THETAP(PADNUM)
DELTA = XX(PADNUM)

!Calculate pad pressure distributions


CALL PRESSCALC(GDOF, EDOF, NBN, NEX, NEZ, NE, NNX, NNZ,

205
* NDOF, HBW, NPE, U, MU, P0, L, PADANGLE, M,OFFSET,
* THP, TP, VX, VY, VD, RJ, RP,ECC, CB, DELTA, THC, PRESS)

!Calculate the x,y force components ON each pad


CALL PLOADCALC(NE, NPE, EDOF, GDOF, NEX, NEZ, NNX, NNZ,
* RJ, THP, OFFSET, PADANGLE, L, PRESS,
* FX(PADNUM), FY(PADNUM))

END DO

!Add up x and y forces on the pads


SFX = SUM(FX)
SFY = SUM(FY)

END IF

!END IF-LOOP FOR PERTURBATIONS


END IF

10 FORMAT (I3, 7(1X, E8.2))

END

!*************************************************************************

206
Appendix C: Matlab Graphics Program for

Fortran TPJB Program Data

%***********************************************************************
%THIS PROGRAM IS THE GRAPHICS FILE FOR THE MORTON_TPJB.FOR FORTAN
%PROGRAM.
%***********************************************************************

%***********************************************************************
% EXTRACT DATA FROM INPUT FILE
%***********************************************************************
% Remove all prior definitions from MATLAB's memory
clear all
close all

%Enter FORTRAN output unbalance file name


fname = input('Please enter Fortran output file name: ','s');
ti = input('Please enter graph title:\n','s');
thrfac = input('What is the threshold fraction of rotor weight (def =
0.15):\n');
if isempty(thrfac) == 1
thrfac = 0.15;
end

dat = load(sprintf('%s',fname));
ldat = length(dat);

% Define graphics input parameters


NPADS = dat(ldat,1); % Number of pads
NS = dat(ldat,2); % Number of speeds
M = dat(ldat,3); % Number of nodes in theta direction
(NNX)
N = dat(ldat,4); % Number of nodes in z direction
(NNZ)

NAVG = dat(ldat-1,1); % Number of temperature averages


Rj = dat(ldat-1,2); % Journal radius (m)
cb = dat(ldat-1,3); % radial brg clearance (m)
W = dat(ldat-1,4); % rotor weight (N)

% Get text coordinates for T,P,H graph


for i = 1:NPADS
j = NS*(NAVG+2) + i;
xt(i) = dat(j,1);
end
yt = dat(ldat,5);
zt = 0.9*dat(ldat,6);

% Extract other data

207
for i = 1:NS
i2 = (i-1)*(NAVG+2) + 1;
i3 = i*(NAVG+2);

%units = m, rad
xamp(i) = dat(i2,1);
xph(i) = dat(i2,2);
yamp(i) = dat(i2,3);
yph(i) = dat(i2,4);
xjeq(i) = dat(i2,5);
yjeq(i) = dat(i2,6);

speed(i) = dat(i3,1);
Uthr(i) = dat(i3,2);
U(i) = dat(i3,3);
DT(i) = dat(i3,4);
phi(i) = dat(i3,5);
eps0(i) = dat(i3,6);

for j = 1:NAVG
j2 = i2 + j;
delT(j,i) = dat(j2,1);
end

end

omega = speed*2*pi/60;

%***********************************************************************
% PROGRAM TO PLOT TEMPERATURE, PRESSURE AND FILM THICKNESS FOR LAST SPD.
%***********************************************************************
% Calculate index for centerline temperatures
if mod(N,2) == 0
ic = N/2;
else
ic = (N+1)/2;
end

%Loop over pads and read in data from appropriate files


for pn = 1:NPADS

switch pn
case 1,
fn = 'PAD1';
case 2,
fn = 'PAD2';
case 3,
fn = 'PAD3';
case 4,
fn = 'PAD4';
case 5,
fn = 'PAD5';
case 6,
fn = 'PAD6';
case 7,
fn = 'PAD7';
case 8,
fn = 'PAD8';
case 9,
fn = 'PAD9';

208
case 10,
fn = 'PAD10';
end

fid = fopen(fn);
A = fscanf( fid, '%g %g %g %g %g', [5 inf]);
A = A';
fclose(fid);

for j = 1:M
r = 1 + (j-1)*N;
r2 = ic + (j-1)*N;
X(j) = A(r,1); % theta (degrees)
TC(j) = A(r2,3); % centerline temperature (deg F)
end

for i = 1:N
Y(i) = A(i,2); % z (inches)
end

for j = 1:M
for i = 1:N
r2 = i + (j-1)*N;
Z1(i,j) = A(r2,3); % temperature (deg F)
Z2(i,j) = A(r2 ,4); % pressure (psi)
Z3(i,j) = A(r2,5); % film thickness (mils)
end
end

figure(1)
subplot(3,1,1)
surf(X, Y, Z1)
view(-10,40)
title(sprintf('%s%s %6.0f %s', ti,': Speed =',...
speed(NS),'RPM'))
text(xt(pn),yt,zt,fn)
zlabel('Temperature (F)')
grid on
hold on

subplot(3,1,2)
surf(X, Y, Z2)
view(-10,40)
zlabel('Pressure (psi)')
grid on
hold on

subplot(3,1,3)
surf(X, Y, Z3)
view(-10,40)
xlabel('Theta (deg.)')
ylabel('Z (in.)')
zlabel('h (mils)')
grid on
hold on

end

209
%***********************************************************************
%PROGRAM TO GENERATE ORBIT PLOTS
%***********************************************************************
%Do calculations in speed loop
for i = 1:NS

%angular speed
w = omega(i);

%orbit parameters
Ax = xamp(i); Ay = yamp(i);
phx = xph(i); phy = yph(i);

%Scale factor for the journal radius in the orbit plots


sf = 0.2*min([Ax Ay])/Rj;

%get coordinates of journal static equilibrium position


xje = xjeq(i);
yje = yjeq(i);

%obtain average parameters for each orbit


for i2 = 1:NAVG
t = i2*2*pi/(w*NAVG);

%plot orbits
figure(1+i)
orbplt(cb,delT(i2,i),t,w,sf,Rj,Ax,Ay,phx,phy,xje,yje);

end

figure(1+i)
title(sprintf('%s%s %7.0f %s %5.2f %s',ti,':',speed(i),...
' RPM, Avg DT = ',abs(DT(i)),' deg F'))
xlabel('x/Cb')
ylabel('y/Cb')
hold on

end

%***********************************************************************
%PROGRAM TO GENERATE UNBALANCE PLOTS AND CALCULATE THRESHOLD SPEEDS
%***********************************************************************
%Plot unbalance curves
figure(NS+2)
sp2 = min(speed):0.001*(max(speed)-min(speed)):max(speed);
om2 = sp2*2*pi/60;
krpm = 1e-3*sp2;
unb = spline(speed,U,sp2);
unbthr2 = (thrfac*W./(om2.^2))*2.2046*16/0.0254;

j = 1;
for i = 1:length(sp2)-1
diff1 = unb(i)-unbthr2(i);
diff2 = unb(i+1)-unbthr2(i+1);
if diff1*diff2 < 0.00 | diff1 == 0.00
Nthr(j) = sp2(i);
Uval(j) = unbthr2(i);
j = j + 1;
end
end

210
plot(krpm,unb,'b--'), hold on

%Threshold unbalance line


fac = thrfac;
unbthr = (fac*W./(om2.^2))*2.2046*16/0.0254;
plot(krpm,unbthr,'r-'), hold on

%Indicate threshold speeds if they exist


if j > 1
for i = 1:length(Nthr)
set(text(Nthr(i)*1e-3,Uval(i),...
sprintf('%7.0f %s',Nthr(i),'RPM')),'Rotation',[90],...
'Color',[0 0 0],'FontSize',[8])
plot(Nthr(i)*1e-3,Uval(i),'ko'), hold on
end
end

%Label plot
legend('U','Uthr')
title(ti)
xlabel('Speed (krpm)')
ylabel('Unbalance (oz in)')

disp(sprintf('%7s %7s %7s %7s %7s','Speed','U','Uthr','Phi','DT'))


disp(sprintf('%7s %7s %7s %7s %7s','(RPM)','(oz in)','(oz
in)','(deg)','(deg F)'))

for i = 1:NS
disp(sprintf('%7.0f %7.2f %7.2f %7.1f
%7.1f',speed(i),U(i),Uthr(i),phi(i),abs(DT(i))))
end

%Display message box


if j > 1 | U(NS) > Uthr(NS)
msgbox(sprintf('%s%4.2f%s','Morton Effect will occur when U > Uthr
(Force = ',thrfac, 'W)'))
else
msgbox('Morton Effect will not occur within the given speed range')
end

%***********************************************************************

%***********************************************************************
% orbplt.m = function used by MORTON_TPJB. This function plots the
%total synchronous orbit and appropriately scales the journal radius to
%show the locations of the cold and the hot spots at different points in
%the orbit.
%***********************************************************************

function [] = orbplt(cb,dT,t,w,sf,Rj,Ax,Ay,phx,phy,xje,yje)

Rj = sf*Rj/cb;

211
Ax = Ax/cb; Ay = Ay/cb;
xje = xje/cb; yje = yje/cb;

g0 = atan2(Ay*sin(phy),Ax*cos(phx));
th = 0:0.01*pi:2*pi;

t1 = 0:0.01*2*pi/w:2*pi/w;
xp = xje + Ax*cos(w*t1 + phx);
yp = yje + Ay*sin(w*t1 + phy);

plot(xp,yp,'k-')
axis equal
hold on

xoj = xje + Ax*cos(w*t + phx);


yoj = yje + Ay*sin(w*t + phy);

xh = xoj + Rj*cos(w*t + g0);


yh = yoj + Rj*sin(w*t + g0);

xc = xoj + Rj*cos(w*t + g0 + pi);


yc = yoj + Rj*sin(w*t + g0 + pi);

xj = xoj + Rj*cos(th);
yj = yoj + Rj*sin(th);

plot(xh,yh,'rx',xc,yc,'bo',[xc xh],[yc yh],'g',...


xj,yj,'k')
axis equal

hold on

%***********************************************************************

212
References

Abdel-Hadi, E.A.A., T.R. Hsu and K.S. Bhatia, “Upwind Finite Element Analysis of
Advection-Diffusion Equation”, Numerical Methods in Thermal Problems, Vol. 5, Pine
Ridge Press, Swansea England, 1985, pp. 756-770.

Allaire, P.E., J.K. Parsell and L.E. Barrett, “A Pad Perturbation Method for the Dynamic
Coefficients of Tiliting-Pad Journal Bearings”, Wear, 72, 1981, pp. 29-44.

American Petroleum Institute, “Centrifugal Compressors for General Refinery Services”,


API standard 617, Fourth Edition, November, 1979.

Cameron, A., The Principles of Lubrication, 1966.

de Jongh, F.M. and P.G. Morton, “The Synchronous Instability of a Compressor Rotor
due to Bearing Journal Differential Heating”, ASME Paper 94-GT-35.

de Jongh, F.M. and P. van der Hoeven, “Application of a Heat Barrier Sleeve to Prevent
Synchronous Rotor Instability”, Proceedings of the 27th Turbomachinery Symposium,
Turbomachinery Laboratory, Texas A&M University, College Station, Texas, 1998.

Dimarogonas, A.D., “Newkirk Effect: Thermally-Induced Dynamic Instability of High-


Speed Rotors”, International Gas Turbine Conference, Washington, D.C., April 1973,
ASME Paper No. 73-GT-26.

Ettles, C.M., “The Analysis of Pivoted Pad Journal Bearing Assemblies Considering
Thermoelastic Deformation and Heat Transfer Effects”, Tribology Transactions, Volume
35, No.1, 1992, pp.156-162.

Faulkner, H.B., W.F. Strong and R.G. Kirk, “Thermally Induced Synchronous Instability
of a Radial Inflow Overhung Turbine, Part II”, Proceedings of ASME Design Engineering
Technical Conferences, September 1997, Sacramento, California, DETC97/VIB-4174.

Goldman, P. and A. Muszynska, “Rotor-to-Stator, Rub-related, Thermal/Mechanical


Effects in Rotating Machinery”, Chaos, Solitons and Fractals, Vol. 5, No. 9, 1995, pp.
1579-1601.

Goldman, P., A. Muszynska and D.E. Bently, “Thermal Bending of the Rotor due to
Rotor-to-Stator Rub”, ISROMAC-7, Honululu, Hawaii, February 1998.

Gomiciaga, R. and P.S. Keogh, “Orbit Induced Journal Temperature Variation in


Hydrodynamic Bearings”, Journal of Tribology, Vol. 121, January 1999, pp. 77-84.

213
Ha, H.C., H.J. Kim, K.W. Kim, “Inlet Pressure Effects on the Thermohydrodynamic
Performance of a Large Tilting Pad Journal Bearing”, Journal of Tribology, Vol. 117,
January 1995, pp.160-165.

Ishida, Y., T. Ikeda, T. Yamamoto, “Transient Vibration of a Rotating Shaft with


Nonlinear Spring Characteristics during Acceleration through a Major Critical Speed”,
JSME International Journal, Vol. 30, No. 261, 1987, pp. 458-466.

Kellenberger, W., “Spiral Vibrations due to the Seal Rings in Turbogenerators: Thermally-
Induced Interaction between Rotor and Stator”, Journal of Mechanical Design, Vol. 102,
January 1980, pp. 177-184.

Keogh, P.S., and P.G. Morton, “The Dynamic Nature of Rotor Thermal Bending due to
Unsteady Lubricant Shearing within a Bearing”, Proceedings of the Royal Society of
London, Series A, Vol. 445, 1994, pp. 273-290.

Kim, J., A.B. Palazzolo, R.K. Gadangi, “TEHD Analysis for Tilting-Pad Journal Bearings
using Upwind Finite Element Method”, Tribology Transactions, Volume 37, 1994, No. 4,
pp. 771-783.

Kirk, R.G., Course Notes for ME 5504 - Introduction to Rotordynamics Analysis, Fall
1998.

Knight, J.D., “Prediction of Temperatures in Tilting Pad Journal Bearings”, Tribology


Transactions, Volume 33, 1990, No. 2, pp. 185-192.

Kroon, R.P. and W.A. Williams, “Spiral Vibration of Rotating Machinery”, Proceedings of
5th International Congress of Applied Mechanics, Wiley, New York, 1939, p. 712.

Larsson, B., “Journal Asymmetric Heating - Part I: Nonstationary Bow”, Journal of


Tribology, Vol. 121, January 1999, pp. 157-163.

Larsson, B., “Journal Asymmetric Heating - Part II: Alteration of Rotor Dynamic
Properties”, Journal of Tribology, Vol. 121, January 1999, pp. 164-168.

Monmousseau, P., M. Fillon, J. Frene, “Transient Thermoelastohydrodynamic Study of


Tilting-Pad Journal Bearings - Comparison between Esperimental Data and Theoretical
Results”, Journal of Tribology, Vol. 119, July 1997, pp. 401-407.

Newkirk, B. L., “Shaft Rubbing”, Mechanical Engineering, Vol. 48, No. 8, August 1926,
pp. 830-832.

Nicholas, J.C., E.J. Gunter and P.E. Allaire, “Stiffness and Damping Coefficients for the
Five-Pad Tilting-Pad Bearing”, ASLE Transactions, Volume 22, 2, 1977, pp. 113-124.

214
Pinkus, O., Thermal Aspects of Fluid Film Tribology, 1990.

Reddy, J.N., An Introduction to the Finite Element Method, 1993.

Schmied, J., “Spiral Vibrations of Rotors”, Rotating Machinery Dynamics, Vol. 2, 1987,
pp. 449-456.

Shapiro, W. and R. Colsher, “Dynamic Characteristics of Fluid-Film Bearings”,


Proceedings of the 6th Turbomachinery Symposium, College Station, Texas, 1977, pp. 39-
53.

Someya, T., Journal-Bearing Databook, 1989

Vance, J.M., Rotordynamics of Turbomachinery, 1988.

Wang, X., “The Influence of a Skewed Disk on a Flexible Rotating Shaft”, M.S. Thesis in
Engineering Mechanics, Virginia Tech, December 1997.

Williams, J.A., Engineering Tribology, 1994.

215
Vita

Avinash Chetnand Balbahadur was born on May 29, 1973 in Georgetown, Guyana. He

graduated from Queen’s College Secondary School (Guyana) in 1991 and subsequently

pursued studies at Dartmouth College in New Hampshire, U.S.A.. In 1995, he received

an A.B. in Engineering Sciences and a B.E. in Chemical Engineering. After obtaining an

M.S. degree in Mechanical Engineering at the same college, he enrolled in the Mechanical

Engineering Ph.D. program at Virginia Tech in 1997. His doctoral research has lead to

the following publications:

A.C. Balbahadur and R.G. Kirk, “A Model for Thermal Bending in Overhung Rotors”,
Proceedings of the ASME Design Engineering Technical Conference, 17th Biennial
Conference on Mechanical Vibration and Noise, September 12-15, 1999, Las Vegas,
Nevada, DETC99/VIB-8298.

R.G. Kirk and A.C. Balbahadur, “Thermal Distortion Synchronous Rotor Instability”,
IMechE 2000, C576/041/2000.

216

You might also like