Текст
                    570/^03 \r.:?>■'<-:.b
Theory of
Hydrodynamic Lubrication
OSCAR PINKUS, M.M.E.
Technical Research Group, Inc.
Syosset, A\Y.
BENO STERNLICHT, PH.D.
Consulting Engineer, General Engineering Laboratory
General Electric Co.
McGRAW-HILL BOOK COMPANY, INC.
New York Toronto London
1901


Engin. Library 1 THEORY OF HYDRODYNAMIC LUBRICATION Copyright © 1961 by the McGraw-Hill Book Company, Inc. Printed in the United States of America. All rights reserved. This book, or parts thereof, may not be reproduced in any form without permission of the publishers. Library of Congress Catalog Card Number: 60-14450. 50050 THE MAPLE PRESS COMPANY, YORK, PA.
PREFACE For a number of years it has been apparent that a need exists for a book on the theory of hydrodynamic lubrication. In response to this need, the Hydrodynamic Research Technical Committee of the Lubrication Division of the American Society of Mechanical Engineers, under the chairmanship of Beno Sternlicht, recommended the writing of this book. The book was written in equal partnership by both authors and with the Committee’s intentions as a guide. We have accordingly attempted to offer here a unified treatment of hydrodynamic lubrication. We have aimed at three objectives. By the application of the general principles of fluid flow to the circumstances of bearing operation, we have stated the differential equations of lubrica¬ tion, including energy and elasticity considerations. We have then pre¬ sented techniques of solving these equations either analytically or by means of analogue and digital computers. And lastly, we have given exact or approximate solutions of the differential equations involved which provide a basis for the design and the solution of specific bearing problems. The presentation of lubrication theory and the solution of bearing problems involve a good deal of mathematics. We have, however, used no more than a minimum of space for discussing the mathematics involved. This was done advisedly. We have used mathematics as a tool for the presentation of lubrication theory and have devoted our main effort to the presentation. We have for the same reasons omitted many of the intermediate steps whenever we considered the mathematical operations to be sufficiently straightforward or standard. We have, however, been careful to state the assumptions and simplifications underlying the final expressions. The subject matter can be broken up into several groups. Chapter 1 lays the mathematical foundations of the material and derives the differential equations of lubrication in their most general form. Chapter 2 deals with simple configurations which are a part of or constitute models of bearings and bearing systems. Chapters 3 to 13 deal specifically with bearings. Perhaps the inclusion of hydrostatic bearings may seem inconsistent with the title of the book; however, both functionally and
vi Preface mathematically, hydrostatic bearings are closely linked to the general subject of hydrodynamic lubrication. The chapter on the Extension of the Classical Theory is an attempt to remove some of the restrictions inherent in the Reynolds equation with the object of finding solutions applicable to such subjects as thick film lubrication and parallel sliders, subjects for which the Reynolds equation collapses or does not hold. Chapter 15 is the only one dealing with experiments. This chapter is intended, not to offer a record of test data, but rather to portray through experimental evidence the validity of the basic propositions of hydrodynamic lubrication. This book draws its material from many sources as well as the authors’ published and unpublished work. We have avoided using proper names except where they are a part of the technical vocabulary. Each chapter is followed by a list of sources, and we have used the term “sources” deliberately to emphasize that the particular chapter is derived from them. The listings at the end of each chapter constitute the acknowledg¬ ment due to the individual contributions. We have not offered a list of references, for the material is too extensive; bibliographies can be found elsewhere. The authors wish to express their appreciation to the Research Tech¬ nical Committee of the Lubrication Division of the ASME for its sponsor¬ ship and interest and to the General Electric Company for the use of its facilities in preparing this book. 0. Pinkus B. Sternlicht
CONTENTS Preface v Nomenclature xii 1. Basic Differential Equations 1 1-1. The Navier-Stokes Equations 1 1-2. The Generalized Reynolds Equation 5 1-3. Flow and Shear Equations 12 1-4. Derivation of Energy Equation 14 1-5. Equation of State 22 2. Hydrodynamics of Simple Configurations 24 General Equations of Motion for Compressible Fluids 24 2-1. The Theoretical Equation for Laminar Flow 25 2-2. The Empirical Equation for Turbulent Flow 25 Flow through Narrow Slots 27 2-3. Isothermal Flow 27 Constant-area Slot 27 Diverging Width 27 Diverging Depth 28 2-4. Flow through Orifices in Scries 28 Incompressible Flow 31 2-5. Flow between Parallel Walls 31 2-6. Circumferential Flow between Concentric Cylinders 32 2-7. Axial Flow in Cylinders 34 Concentric Cylinders 34 Flow through Eccentric Cylinders 35 3. Incompressible Lubrication; One-dimensional Bearings 37 The Real Bearing 37 One-dimensional Journal Bearings 41 3-1. Infinitely Long Bearing 42 3-2. Infinitely Short Bearing 48 3-3. Partial Bearings 50 3-4. Fitted Bearings 51 3-5. Floating-ring Bearings 53 3-6. Porous Bearings 54 One-dimensional Thrust Bearings 56 3-7. Plane Sliders 58 3-8. Curved Sliders 59 vii
viii Contents 3-9. Step Bearings 60 3-10. Composite Bearings 62 3-11. Pivoted-shoe Bearings 64 4. Incompressible Lubrication; Finite Bearings 68 Finite Journal Bearings 68 Analytical Methods 69 4-1. Journal Bearing with Axial Feeding 71 4-2. Journal Bearing with Circumferential Feeding 75 Numerical Methods 79 4-3. Digital Computers 79 4-4. Electrical Analogues 81 Journal Bearing Solutions 85 4-5. Full Journal Bearings 85 4-6. Centrally Loaded Partial Bearings 92 4-7. Eccentrically Loaded Partial Bearings 93 4-8. Axial-groove Bearings 93 4-9. Noncircular Bearings Ill Finite Thrust Bearings 116 Analytical Solutions 116 4-10. The Step Bearing 118 4-11. Slider with Exponential Film Shape ' 122 Numerical Solutions 124 4-12. Slider Bearing; Semianalytical Methods 124 4-13. Sector Pad; Computer Solution 129 6. Hydrodynamic Gas Bearings 136 General Considerations 136 Limiting Characteristics 138 Infinitely Long Slider Bearings 139 5-1. Parallel Surface 139 5-2. Plane Inclined Slider 141 5-3. Composite Slider 145 5-4. Step Slider 148 5-5. Convergent-Divergent Slider 149 5-6. Curved Slider 151 Finite Slider Bearings 152 5-7. Plane Inclined Slider 152 Infinitely Long Journal Bearings 152 5-8. Journal Bearing with Inertia Considered 152 5-9. Journal Bearing with Inertia Neglected 157 f5-10. Numerical Solution 158 *5-11. Katto and Soda Solution 159 Finite Journal Bearings 163 5-12. Perturbation Solution 163 5-13. Linearized ph Solution 168 5-14. Numerical Solution 171 6. Hydrostatic Bearings 177 Plain Journal Bearings 178
Contents ix 6-1. Incompressible Lubrication 178 Laminar Feeding 178 Turbulent Feeding 182 Rotational Considerations 187 6-2. Compressible Lubrication 190 Laminar Feeding 190 Turbulent Feeding 193 Static and Dynamic Characteristics of Gas Journal Bearings .... 196 Step Thrust Bearing 200 Isothermal Operation 200 6-3. Compressible Lubrication 200 6-4. Incompressible Lubrication 203 Adiabatic Operation 204 6-5. Incompressible Lubrication 205 6-6. Compressible Lubrication 206 Self-excited Vibrations in Gas-lubricated Step Thrust Bearing .... 207 7. Squeeze Film and Dynamic Loading 213 Dynamically Loaded Bearings 213 The Reynolds Equation for Dynamic Loading 214 Cyclic Squeeze Films in Journal Bearings 217 7-1. Constant Loads 219 7-2. Alternating Loads 219 7-3. Rotating Loads 219 Noncyclic Squeeze Films 220 7-4. Journal Bearings 221 7-5. Spherical Bearings 222 7-6. Conical Seats 223 7-7. Sliders and Rectangular Plates 223 7-8. Elliptical and Circular Plates 225 7-9. Miscellaneous Configurations 226 Dynamic Loading of Journal Bearings 227 7-10. Constant Unidirectional Loads 227 7-11. Variable Unidirectional Loads 229 Sinusoidal Loading 231 Square-wave Loading 233 7-12. Constant Rotating Loads 235 7-13. Variable Rotating Loads 237 Dynamic Loading of Journal Bearings with No Negative Pressures. 238 7-14. Solutions for Prescribed Loci 242 7t15. Solutions for Prescribed Loads 247 Solutions for Finite Journal Bearings 254 7-16. Dynamic Loading 254 7-17. Squeeze Films 260 8. Hydrodynamic Instability 264 8-1. The Mechanics of Hydrodynamic Instability 264 8-2. Hydrodynamic Forces on Journal 265 8-3. Threshold for Half-frequency Whirl 269 8-4. Forced Vibration of Vertical Rotor 270 8-5. Fluid-film and Rotor Resonance 272
X Contents 8-6. Equations of Small Oscillations 274 8-7. Equations of Motion for Large Displacements 281 9. Adiabatic Solutions 286 Introduction 286 The Thermal Wedge 288 One-dimensional Solutions 292 9-1. Parallel Slider with p = f(T), p = f(p,T) 292 9-2. Step Slider with p = f(x), p = f(p,T) 294 9-3. Exponential Slider with p = f(p,T) 297 Finite Solutions 298 10. Elasticity Considerations 306 Introduction 306 One-dimensional Solutions 306 10-1. The Perfectly Elastic Journal Bearing 306 10-2. Spring-supported Thrust Bearing 308 10-3. Pivoted Shoe with Elastic Deformation 313 Two-dimensional Solutions of Centrally Pivoted Sectors 315 11. Hydrodynamics of Rolling Elements 328 General Remarks 328 Fluid Film with Rigid Surfaces 329 11-1. Solutions with Constant Viscosity 329 11-2. Viscosity as a Function of Pressure 335 Fluid Film with Elastic Deformation 336 11-3. Solution with Constant Viscosity 338 11-4. Viscosity as a Function of Pressure 339 12. Inertia and Turbulence Effects 351 Introduction 351 Effects of Fluid Inertia 352 Significance of Inertia Terms 352 Iteration Method 354 12-1. Slider with Inertia Considered 355 12-2. Journal Bearing with Inertia Considered 357 Method of Averaged Inertia 360 12-3. Squeeze Films 362 12-4. Journal Bearing 362 12-5. Slider Bearing 364 Acceleration Effects in Bearings 365 Effects of Turbulence 367 12-6. Criteria of Fluid Instability 368 12-7. Turbulent Operation of Journal Bearings 371 12-8. Turbulent Operation of Slider Bearings 373 13. Non-Newtonian Fluids 380 General Remarks 380 Bingham Plastics (Greases) as Lubricants 381 13-1. Rheodynamic Bearings 381 13-2. Squeeze Films 385
Contents xi 13-3. Rheostatic Bearings 387 Viscoelastic Lubricants 396 14. Extension of the Classical Theory 403 The Restrictions of Lubrication Theory 403 The Inflow Wedge 404 14-1. The Case of Parallel Sliders 404 14-2. The General Case of Plane Sliders 409 Variations across the Fluid Film 416 14-3. Sliders with High Angle of Inclination 415 14-4. Journal Bearings with Large Clearances 419 16. Experimental Evidence 426 Pressure Profiles 427 15-1. Liquid Lubricants 427 15-2. Gaseous Lubricants 428 15-3. Grease as a Lubricant 428 The Fluid Film; Cavitation 432 15-4. The Fluid Film under Steady Loading 433 15-5. The Fluid Film under Dynamic Loading 439 Locus of Shaft Center 441 15-6. Steady Loading 442 15-7. Dynamic Loading 443 15-8. Instability 450 Turbulence 452 15-9. Breakdown of Laminar Flow 452 15-10. Effect on Bearing Performance 456 A’ame Index 459 Subject Index 461
NOMENCLATURE Unless otherwise specified, the following symbols are used in the text: B breadth, width (parallel to direction of motion) C radial clearance D diameter E energy, Young’s modulus F frictional force, force G weight flow rate H power, work rate I moment of inertia J mechanical equivalent of heat K spring constant L bearing length (normal to direction of motion) M mass, moment N revolutions per unit time P unit loading = W/LD Q volume flow rate R bearing radius (R perfect-gas constant Re Reynolds number S Sommerfeld number = (i*N/P)(R/C)2 T temperature U linear velocity V velocity, volume W load a ratio of inlet to outlet film thickness b damping coefficient c specific heat e eccentricity f coefficient of friction, dimensionless force = l/S 9 constant of gravitational acceleration h film thickness k ratio of specific heats, thermal conductivity, spring constant m mass xii
Nomenclature xiii n polytropic constant p pressure q heat flow rate, volume flow rate per unit length q dimensionless flow coefficient = Q/tRCNL t time u,v,w linear velocity components 2 center of pressure x,?/,z rectangular coordinates a dimensionless taper, load angle 0 angular span of bearing arc or sector 6 amount of taper = (h\ — h2), ellipticity ratio € eccentricity ratio = e/C A bearing number = §nUB/pahl or (6fjLO)/pa)(R/C)2 M absolute viscosity v kinematic viscosity p density r shear stress <p attitude angle o) angular velocity Subscripts H horizontal L laminar, load P load R radial T tangential, turbulent V vertical, volume a ambient b bearing c common, critical j journal 1 laminar r radial s slider, supply t tangential, turbulent avg average max maximum min minimum opt optimum red reduced 0 pertains to point of maximum pressure
Nomenclature beginning, inlet end, outlet Superscripts dimensionless (x = x/B) first derivative with respect to time second derivative with respect to time third derivative with respect to time
CHAPTER 1 BASIC DIFFERENTIAL EQUATIONS The study of hydrodynamic lubrication is, from a mathematical stand¬ point, the study of a particular form of the Navier-Stokes equations. This particular differential equation was formulated by Osborne Reynolds in 1886 in the wake of a classical experiment by Beauchamp Tower in which the formation of a thin fluid film was for the first time observed and understood to be the basic mechanism of hydrodynamic lubrication. This Reynolds equation can be deduced either from the Navier-Stokes equations or from first principles, provided, of course, that the same assumptions are used in both methods. The Reynolds equation contains viscosity, density, and film thickness as parameters. These parameters both determine and depend on the temperature and pressure fields and on the elastic behavior of the bearing surfaces. Thus, to get a complete and accurate representation of the hydrodynamics of the lubricating film, it is oftentimes necessary to consider simultaneously the Reynolds equa¬ tion, the energy equation, the elasticity equation, and the equation of state. This chapter deals with the mathematical formulation of these equations and serves as a basis for the solution of bearing problems in subsequent chapters. We shall first derive, in as simple a manner as possible, the Navier-Stokes equations and subsequently reduce them to the Reynolds equation, thereby showing the restrictions and assumptions inherent in the equations used in the solution of lubrication problems. 1-1. The Navier-Stokes Equations. In a viscous fluid, across each of three mutually perpendicular surfaces, there are three stresses, giving a total of nine stress components. Figure 1-1 shows the components that act across a surface perpendicular to the z axis. The first subscript of each stress is z to indicate this fact. The second subscript designates the axis parallel to which the stress acts. There are three similar stresses for the top face of the cube and three more for the right face. To have equilibrium among the forces acting on the cube, the stresses must be symmetric, i.e., the subscripts can be reversed in order, viz., Txy ~ Tyx Txz ~ Tzx TyZ — TZy ( These three stress components all act tangentially to the surface across which they are propagated, and they are the shear components. In
2 Theory of Hydrodynamic Lubrication addition the pressure p of the fluid is often considered to be the average of the three normal components, which act normally to their surfaces; i.e., TXX + Tyy + TtZ = — 3 P The minus sign is used because negative pressures are compressive, whereas positive stresses are tensile. The magnitude of the stresses depends on the rate at which the fluid is being distorted. For most fluids the de¬ pendence is of the form (dui . duA to, + dxj~ iiiV in which the u1 s are the components of the velocity vector. The term in parentheses measures the distortion, as opposed to rigid motion, of the fluid. This equation indicates that the stress components are proportional to the distortion of the fluid and that the constant of proportionality is y, the viscosity. From the normal stress compo¬ nents—the three components in which the two subscripts are the same— one must subtract an additional stress due to the static pressure of the fluid. 8ij is the Kronecker delta to indicate that p is to be subtracted only in case the two subscripts are equal. For the three shear stress components (du . dv\ (du . dw\ (dv , dw\ = -, + Tx) = + T"~M^ + 5ir; in which u} v, and w are the x, y, and z components of velocity. The other three stress components are the normal components, and for them, i o du . o dv . o dw rxx=-p + 2»-^ ryw=-p + 2„- T„=-p + 2» — The sum of these normal components is r„ + r„ + r„ = -3p + 2, (g + g + g) = —3p 4- 2yd The term in parentheses is the divergence of the velocity vector, the dilatation. It measures the rate at which fluid is flowing out from each point; i.e., it measures the expansion of the fluid. For brevity, it can be indicated by 6. In order for the three normal components to add to — 3p, some multiple of 0 must be added to them; they must be redefined as Fig. 1-1. Stresses on one face of fluid element.
Basic Differential Equations 3 <rx = — p -f X0 + 2/z — au — — p + X0 + 2/i — (Tz = — p + X0 + 2/i-T- The coefficient of 0 is X, at present an undetermined quantity. The sum of these normal components is —3p provided that Some fluids have a “volume viscosity” that measures their resistance to volume changes, just as conventional viscosity measures the resistance to flow. In the case of a fluid with a volume viscosity, X + is not zero. For the time being, both X and p can be retained in the equations; subsequently the volume viscosity can be assumed to be zero and X can be expressed in terms of p. The three components of acceleration of the fluid are the three total derivatives Du/dt, Dv/dtf and Dw/dt. The mass of an element of fluid with dimensions dx, dy, and dz is p dx dy dz; thus the components of the force required to accelerate the element are Total derivatives of the velocity components are calculated by treating the velocity as a function of x} y, z, and t in which x, y, and z are them¬ selves functions of t. The partial derivatives of x} y, and z with respect to time are, of course, the velocity components w, v, and w. The total time derivative measures the change in velocity of one particle of fluid as it moves about in space; the partial time derivative measures the change in velocity of whatever particles of fluid occupy one particular location. The forces needed to accelerate the element of fluid are supplied by an external force field, perhaps gravity, and by pressure or stress gradients within the fluid. If the components of the external force field per unit mass are X, Y, and Z, these forces are equal to Xp dx dy dz, Ypdxdy dz, Zp dx dy dz. The forces due to the stress gradients must be added to the external force. Three stresses tend to move the element in the x direction. X = p dx dy dz p-^dx dy dz p dx dy dz Du diL fir. fin fin Ha. fiz du dt
Theory of Hydrodynamic Lubrication Thus, a change in tXV) for example, across the cube and through a distance dy. is drXi dy dy This stress acts on the face of a cube with area dx dz and produces a force: Force = dy dx dz dy There are similar expressions for rIX and rxz. When these stress-gradient forces are added to the external force and the common factor dx dy dz is eliminated, Du dt v- i d(Tx . drxu drxz pX + + ~dy + Also, there are similar expressions for the stress gradients that tend to move the element in the y and z directions: V _L dT*u I d<Tu , dr^ p dT = pl + + lij + Tz Dw dr„ dr„z d(Tz PW = pz + ^ + ~ + dx dy dz In case the volume viscosity X + %p is zero, X can be written in terms of p, and we have after replacing m by their proper expressions Du dp d i dx dy "W - pY-?,+§j\r ; .+£ •y-rf-g + sl- +1 [2»? 2fc + *+^')l| dx S\dx dy dz J J) (du dlA] d \ (Sw duN] dv _ 2 (du dv_ dw\ 1) dij ~ 3 \di + d7j “*■ dz)J) p(I+S!)]+£[m(S+£)] dw _ 2 (du dv ) ~dz ~ 3 V^r + dy + (dw . d f (dv dwAl ^ + 3^J + ^K*+-WJ (l-la) (1-16) (1-lc) which are the general Navier-Stokes equations. Equations (1-1) con¬ tain the four unknowns u, v, w, and p. A fourth equation is supplied by the continuity equation:
Basic Differential Equations 5 where m accounts for the presence of sources and sinks. With no sources or sinks present and with the state of the lubricant independent of time, the continuity equation reads + dl^l = o (1-2 a) dx dy dz Equations (1-1) and (1-2) contain the density and viscosity as variable parameters. Thus in order to define fully the problem, functional rela¬ tionships for p and y must be available in the form p = p(p,T) y = y(p,T) In cylindrical coordinates, the’ Navier-Stokes equations read / Du uv\ 1 dp 1 | r 2 dw _ 2/1 drv 1 du P \ dt r ) r dd r dQ (M [r ^0 \r dr r dB + £)]|+'l['‘0S+5)]'+f,['OS <>-"0 (Dv u2\ D dp . d \ \ ndv 2/1 drv . 1 du . dw\l) = + ^ Hr^ + ra5 +ai-JJj + I AT /I*? 4_ On _ A\1 , d_ r dd \r dd + dr r ) J + dz + 7 (dv . ditAl + «F/J dv 1 du vl /t t n Tr-ree-'r\ <Me) d ( dw 2(\drv I du dw\l| . l a r /dv , , i a r (\aw , au\] + r 7r [Mr +*)\ + 7 ae |_p ae + Tz)\ (M/) Dw dp p-W = z~al + with the continuity equation given by d(pu) a(prv) d(pw) = ol ae dr dz 1-2. The Generalized Reynolds Equation. The differential equation originally derived by Reynolds is restricted to incompressible fluids. This, however, is an unnecessary restriction; for the equation can be formulated broadly enough to include effects of compressibility and dynamic loading. We have called this the generalized Reynolds equa¬ tion. At the end of this section we have written down the various forms of this equation as they apply to particular cases of bearing operation.
6 Theory of Hydrodynamic Lubrication The assumptions involved in reducing Eqs. (1-1) to the Reynolds equa¬ tion are, referring to the fluid film of Fig. 1-2, as follows: 1. The height of the fluid film y is very small compared to the span and length x, z. This permits us to ignore the curvature of the fluid film, such as in the case of journal bearings, and to replace rotational by trans¬ lational velocities. 2. No variation of pressure across the fluid film. Thus ? = ° dy 3. The flow is laminar; no vortex flow and no turbulence occur any¬ where in the film. 4. No external forces act on the film. Thus X = Y = Z = 0 5. Fluid inertia is small compared to the viscous shear. These inertia forces consist of acceleration of the fluid, centrifugal forces acting in curved films, and fluid gravity. Thus Du _ Dv _ Dw _ „ dt dt dt 6. No slip at the bearing surfaces. Fio. 1-2. The fluid film. 7- Compared with the two veloc- ity gradients du/dy and dw/dy, all other velocity gradients are considered negligible. Since u and, to a lesser degree, w are the predominant velocities and y is a dimension much smaller than either x or z, the above assumption is valid. The two velocity gradients du/dy and dw/dy can be considered shears, while all others are acceleration terms, and the simplification is also in line with assumption 5. Thus any derivatives of terms other than du/dy and dw/dy will be of a much higher order and negligible. We can thus omit all derivatives with the exception of d2u/dy2 and d2w/dy2. Assumptions 1 to 7 used in Eqs. (1-1) yield 1 dp _ d2u iai-w ( } 1 dp _ d2W . -pTz~W ( 36) Equations (1-3) can also be derived directly by setting up a force balance of an element of fluid in the lubricant film. With external forces and inertia neglected and with no pressure gradients in the y direction, all shear and pressure forces are as given in Fig. 1-3. By summing the forces in the x and z direction, we obtain
Basic Differential Equations 7 ^r* + ^ dz'j dx dy — tx dx dy + (rx + ^ dy^j dx dz — r* dx dz + (p-licdx)dydz-(p + lZdx)dydz = 0 H- ^ drr^ dy dz — tm dy dz + ^rf + ^ dy^j dx dz — r* dx dz + (p~ I2dz)dxdy ~ {v + \idz)dxdy = 0 By canceling like terms and simplifying, drx drx _ dp dy dz dx dr* . drz _ dp dy dx dz We now introduce the assumption of a Newtonian lubricant, i.e., a fluid in which the shear stress is proportional to the rate of shear, or du T* = 11 dy dw tz = y-r~ dy (l-4a) (1-46) It will be recalled that this assumption was made in deriving the Navier-Stokes equations, and it was thus unnecessary to introduce it in dXr (i> + -jjdy)dxdz {rx+ — dz)dxdy 1/ i,d,dy ~~J / / •«- T:dydz 2Txdx^y(iz r/dxdz Txdxdz 1 / ¥{p-t^d!)dxdy (r!^^dy)dydi (j) Forces in / direction {t>) Forces in / direction Fig. 1-3. Forces acting on a fluid element. deriving Eqs. (1-3). Here, however, starting from a basic momentum equation, we must rely on expressions (1-4) to provide us with a relation between velocity and stress. Physically, assumption (1-4) implies that the viscosity is independent of the rate of shear, a phenomenon which is not true of all fluids. By using (1-4) in the preceding equation, we obtain d2u d2u _ 1 dp dy2 dy dz y dx d2w d2yy _ 1 dp dy2 dy dz y dz
8 Theory of Hydrodynamic Lubrication Now, by making use of assumption 7, the second terms on the left-hand side drop out and we have Eqs. (l-3a) and (1-36) derived above: d2u I dp n o i lWal d^ = -f [1-36] dy- y dz By integrating Eq. (l-3a) twice with the boundary conditions, u — U\ at y = 0 u — L\ at y = h we have u = ^ y{y - h) + h -^-y- Ux + | Ui (l-5a) By integrating Eq. (1-36) with the boundary conditions, w = 0 at y — 0 and at y = h which implies a motion of the bearing surfaces only in the x direction, we have w = TMy(y~h) (1'56) Equations (1-5) give the velocity profile in the fluid film as affected by the viscosity y, film shape 6, surface velocities Ux and U2, and the pressure gradients. We now make use of the continuity equation (1-2). The equation can be written d(pv) _ d(pu) d(pw) dy dx dz By replacing u and w by their values from Eqs. (1-5), we have By integrating with respect to y with the conditions v = V at y = 0 and v = 0 at y = 6, we have pF = f0ki[-,ty{y-h)](lu + fo y(y~h)]dy] ~/0 hp\^irUi + lUi\du The upper limit h in the last equation is a function of the coordinates xy z. By making use of the relation
Basic Differential Equations 9 fh(a) d d fh<a> dh(a) Jo dad}J = da Jo f('J'a) d,J ~ ma)'a] we can perform the integration before differentiating to obtain + £(^Y| 2 [dx \6y dx/ dz\bydz/] IT ld(f'r2+f/l) , (1J JJ X d(ph) 1 “ 2 [* di + {Ul ~ Ui) ~dx~J + 6p6— (Ux + U2) ■+■ 12pF (l-6a) The second term on the right-hand side implies a variation of tangential velocities. For steady loads this means a rubber-like stretching of the bearing material. This phenomenon, if at all possible, is a rarity in bearing operation. In dynamic loading, when a radial component of velocity V exists, the tangential velocities may vary. Referring to Fig. 7-1, we can write for a journal bearing the following: U = Rwi + —rp s*n 0 — Ce —1^ — cos 0 at at Tr ~ de n . ~ d((3 -f- y) . _ V = C -r cos 0 + Cc -- —- sin 0 . at at By differentiating U with respect to x, we have dU = IBU dx R d0 1 T„de n „ d(fi + y) • J 1 = R[CdtC0S6 + C(—di~Sm *] = * Thus when dU/dx is replaced by (1 /R)Vt we have for the right-hand side of the Reynolds equation Ciph (tri + r2) + VlpV « 12 ^ V + 12pV ox It - 12p(l+4V Now h/R is a quantity very much smaller than 1, usually of the order of 0.001, and it can be discarded. Thus the final form of the generalized Reynolds equation is d(ph) dx
10 Theory of Hydrodynamic Lubrication The first right-hand term 6(t/i — U2) d(ph)/dx, is obviously the contri¬ bution of the bearing velocities along the oil film, while the term 12pV is due to the relative velocity of bearing surfaces in a direction normal to the fluid film. It is of interest to note that the effect of the term 6(C71 — U2) d(ph)/dx depends on whether the bearing surfaces have trans¬ lational or angular velocities. For a thrust slider if Ui — U2, the first right-hand term of Eq. (1-66) disappears and—since in the absence of any normal movement of bearing surfaces, V = 0—such a bearing has zero load capacity; conversely, if Ui = — U2, the load capacity is doubled. However, in a journal bearing if Ui = t/2, the first right-hand term dis¬ appears, but the angular motion of the two surfaces introduces both Fiq. 1-4. The effect of movement of bearing surfaces. tangential and normal components of velocity. Referring to Fig. 1-4, these velocities are Tangential velocity = U cos a « U dh Normal velocity = U sin a « U tan a = U — = V ox Thus the right-hand side of Eq. (1-66) becomes 2 U dh/dx and the load capacity is doubled; conversely, if the bearing surface rotates in the opposite direction, i.e., if Ui = —1/2, the first right-hand term becomes 2U dh/dx, but then the sign of V is reversed, giving —2U dh/dx, and the net result is zero load capacity. In journal bearings, therefore, when both bearing and journal rotate in the same direction, the velocities are additive; in thrust bearings they are subtractive. This can also be seen intuitively by realizing that in thrust bearings, when both surfaces move with the same velocity, any two points on the opposing surfaces remain at a fixed distance from
Basic Differential Equations 11 each other; in a journal bearing any two points on journal and bearing will approach each other at a rate depending on how fast both of the surfaces move. However, it can be shown that, if the center of curvature of the journal bearing surface does not coincide with the center of rotation, counterrotation at identical speeds will produce hydrodynamic forces. Thus bearings with a noncircular cross section will yield a load capacity even under the above conditions7. Equation (1-66) holds for both compressible and incompressible lubri¬ cants. By setting p = const, the Reynolds equation for incompressible fluids is obtained: £(7!)+ £(7 + In Eq. (1-66) the viscosity p is still treated as a variable, being a function of both the x and z coordinates. The film thickness h, too, is general enough and can be a function of both coordinates. Equations (1-66) and (1-7) are nonhomogeneous partial differential equations of two variables. They are difficult equations to solve, and the degree of complexity depends on the form of the parametric functions p, p, and h and on the boundary conditions. Even for the simplest case of p = p = const and V = 0 when Eq/ (1-66) reduces to + <‘-8> closed solutions are difficult to obtain. Some successful attempts in solving Eq. (1-8) for simple functions h(x) and numerical solutions for the more complicated cases are treated in later chapters. In cylindrical coordinates using the substitutions x — r cos 0 and z = r sin 0 the generalized Reynolds equation becomes IC? %) +; f. (f 8) - «v' - v* ™ <■-») In deriving Eq. (1-66) we have used the boundary conditions in a manner such that V represents the resultant normal velocity regardless of what is instrumental in producing this radial motion. In thrust bearings, the velocity V can come only from the actual normal movement of the sliding surfaces. However, in journal bearings, as we have seen above, a normal relative velocity can come from two sources: from the rotational velocity of the sliding surfaces and also from any actual motion of journal center. It is convenient to have V represent only the radial velocity that results from the motion of shaft center, and we shall thus
12 Theory of Hydrodynamic Lubrication rewrite Eq. (1-66) for journal bearings in the following manner: In most practical cases, the bearing is stationary and only the runner in thrust bearings and the shaft in journal bearings are moving. In that case, Eqs. (1-10) and (1-66) reduce to which is the same for both thrust and journal bearings with U the sliding velocity of either runner or journal. For steady loading (Fo = 0) and incompressible lubricants (p = const) Eq. (1-11) becomes which is the most commonly encountered form of the Reynolds equation. 1-3. Flow and Shear Equations. Several important expressions have been formulated in the process of deriving the Reynolds equation. These are the flow and shear equations of lubrication. It was pointed out that Eqs. (1-5) represent the velocity components of the lubricant in the x and z directions. These equations, when integrated between the two bearing surfaces, provide the lubricant flow at any given section: where F0 now represents the motion of journal center. (1-12) /:[ These integrations yield h3 dp h V2fx dx 2 hz dp 12p dz (l-13a) (1-136) The flow in the z direction will be positive or negative depending on the
Basic Differential Equations 13 sign of the pressure gradients. The flow in the x direction is made up of two components: the pressure flow (h3/\2n)(dp/dx) and the shear flow h(Ui + U2)/2. Its direction will depend on both the magnitude of Ui and Ui and the sign of dp/dx. In polar coordinates Eqs. (1-13) are _ h3 dp , r(o»i -f- o)i)h qe \2ritdB H 2 ~ _ __ h3 dp ^r 12n dr The shear stress from the definition of a Newtonian fluid as given by Eqs. (1-4) is du dw By differentiating Eqs. (1-5) with respect to y, we have Ti = IS(2i/ _ h)+i{U* ~Ui) (1*14a) The value of the shear stress depends on y, and the sections of interest are the two bearing surfaces. Thus at the surface moving with velocity U i we have y = 0 and T*= + Ul) (1'15a) Thus _ _ h dp 2 dz Tz = ~ T^TT- (1-155) At the surface moving with velocity U2 we have y = h and + !"'• (|-10“> '• = \1 <-'-m Since the total force is given by integrating r over the bearing surface, we have F = JJV dA Since Fz = JfrzdA is at right angles to the displacement of the bearing surfaces, the total drag exerted by the moving bearing surface at y = 0
14 Theory of Hydrodynamic Lubrication dx dz (1-17) or y = h is given by IN In polar coordinates the above equation is by expressing Eq. (1-17) in terms of torque rather than force, M h dp y.r(o)2 — o) ± 2r ae + r2 dd dr (1-18) 1-4. Derivation of Energy Equation. In rigorous bearing analysis the variation of viscosity with temperature must be considered. As the fluid is sheared, work is being done on it and there is a temperature rise which in turn reduces the viscosity of incompressible fluids and raises the viscosity of compressible fluids. This variation of viscosity must be Fig. 1-5. Incremental volumes. included in the solution of Reynolds equation. Likewise, from the stand¬ point of heat transfer and thermal distortion, it is desirable to determine the temperature gradients that exist in the bearing. This section deals with the derivation of energy relations which describe the temperature variation in the fluid film. It is desirable to have the modified energy equation in such a form that all variations with y are integrated “out of the picture,” some¬ what analogous to the form of the Reynolds equation. There are two approaches in deriving the modified energy equation: one is to sum ener¬ gies on an incremental control volume of finite height h, as in Fig. l-5a; the second is to sum energies on an incremental control volume of incremental height, as in Fig. 1-56, and then integrate over the height h. Since there is some confusion over what constitutes mechanical energy for a bounded finite-height incremental volume, the second of the two approaches will be used. However, care must be exercised in integrating the boundaries having slope. If certain terms are neglected in the integrand, an erroneous partial differential equation will result.
Basic Differential Equations 15 Control volumes such as illustrated in Fig. 1-5 are imaginary volumes generally fixed in space through which the fluid at continuously varying velocity, temperature, pressure, density, and viscosity is allowed to pass. Since an energy equation is desired, all the component energies will be summed over this volume for a unit interval of time according to the first law of thermodynamics: Ei + Hdo — E, + E0 + Ha, where Ei = energy transported into the control volume E0 = energy transported out of the control volume E, — energy stored transiently in the control volume Hdo ~ work done on the fluid volume by the surroundings Hdb = work done by the fluid volume on the surroundings Steady-state conditions are assumed, so that the above equation becomes E0 — Ei = Hdo — Hdb (1-19) There are two modes in which energy may be transported into and out of control volumes: by conduction according to Fourier’s law and by convection of intrinsic energy, i.e., transport of fluid possessing kinetic energy and internal energy. A possible third mode, radiation, is neglected. The other energies involved in the energy balance are the mechanical works done by the surface stresses and body forces through an incremental distance in an increment of time. For the lubrication problem at hand, body forces, such as gravity, are neglected. The transported energies and mechanical works involved are indicated separately in the control volumes of Fig. l-6a and 6. So as not to encum¬ ber the sketches, not all component energies are indicated. It is to be noted that differential changes in energies are taken about the mid-point 0 in the control volumes. The transported energies of Fig. l-6a summed over the surfaces of the control volume according to the left-hand side of Eq. (1-19) are *• - * - (pfc*+:V+“E*] -' [£(*£)■+ «(*S) where the intrinsic energy is given by e = *±4±^ + Jc.T The mechanical works indicated in Fig. 1-66 must also be summed over the volume surfaces. However, an interpretation of what is meant
16 Theory of Hydrodynamic Lubrication by work done on and work done by a volume in terms of the surface stresses and fluid velocity is first needed. All the works done by the fluid volume are on the upstream surfaces of the control volume, i.e., where the velocity components are in the opposite direction to the stress dog A/ E--q dA etc do/ A/ dir A/ °> + 17 ~'*+77T dT/X A/ dr/ A/ T"+iJ/ T’*'+IrT . du A/ „JjLki7 > it 2 \/ dr A/ d<rr A/ dK A/ U) Fig. 1-6. Control volumes, (a) Transported energies; (b) mechanical energies. components. With this viewpoint in mind, the right-hand side of Eq. 1-19 becomes d HJo Hdb — (llOz “f" tJTU* "f" dy d\ , (UT V<7 y -j- WT zy) a,) j Ax Ay Az By equating the expressions for E and H according to Eq. (1-19), + ^ {utxz -j- xnuz + W0z] |~d(pue) d(pve) d(pwe) [ dx dy dz
Basic Differential Equations and by rearranging some of the terms. 17 [*S-S+-£]-'[£(*SK(*S)+£(*S)] -*(£+»+£)+*(£+&+&) (&+1+b)+('• B+- S+S)+- (w+ S) (I+Ij)+'"(b + £) <■-“> + “Tyi where the first parenthesis term of Eq. (l-20a) reduces to its stated form because of the continuity equation d(pu) d(pv) d(pw) _ dx ^ dy ^ dz The equilibrium equations of fluid flow for steady-state conditions (dv/dt = 0) and zero body force are given by the expressions preceding Eqs. (1-1), namely / dll du : du\ dox drzy p{UYx + Vd-y + WYz) = -dI + ^ ( dv . dv . di>\ drux p\UTx + Vd-y + Wdi) = l>x ( dw dw dw\ drtz drty , do- p\udl + vdj +wdl) = aT + ^ + aF dTx- dz i doy . dryx ^ dy ^ dz The substitution of the expressions for o and r into the last three equa¬ tions yielded for us previously the Navier-Stokes equations. Substitut¬ ing the last three equations and equations for o and r into Eq. (1-206) yields ( de . de . de\ . I" d /, dT\ . d (, dT\ . d /. dT\ 1 p\uTx + vd-y + wTz) ~ ,/[aiV+ 9v\ dy) + d’z\!fe = P (u d[{u2 + v2 + w2)/2] v d[(u2 + »2_±w2)/2] dx d[(u* + ; dy ^ dz j p\dx^ dy ^ dz From e = (u2 + v2 + w2)/2 + JcvT we have jp r u +v ^p.+w a(c’7’) dx dy dz 1 , [du di> dw + ~dz + (1-21)
18 Theory of Hydrodynamic Lubrication where - * [* (£)‘+’ (£)'+*(£)■- i (£+S+£)' . (du dv\2 . (dv . dw\2 . fdw dw\2l + + ^ + + ^ + a?jJ It is to be noted that only a d{cvT)/dt need be added within the first bracket of Eq. (1-21) to make the energy equation applicable also to tran¬ sient states, subject only to the limitation that the flow be laminar. The first bracket term in Eq. (1-21) is the convection of internal energy of the fluid. The second bracket term is the rate of work done by a differential volume of fluid in expansion against the surrounding pressure. fro dh. dh -a Fig. 1-7. Energies at solid boundaries with slope, (a) Transported energies; (6) conducted heat. The third bracket term is the rate of heat conduction in the fluid. And the fourth term is the rate at which kinetic energy is dissipated into heat. A convenient starting point in integration of the energy equation across the film thickness is at Eq. (l-20a). Thus, Eq. (l-20a) becomes /:[ d(pue) d(pwe) 1 , |A =/:[ d d (iKTz + VTvx + WTZX) -f- ~ (UT -1- VTyx + w<Tt) dy “1“ (WTXy “I- V(Ty -f" WTZy) For the moment the terms of interest in the above equation are: pve kd-T dy (UTzy + V<Ty + WTzy) which must be investigated at an incremental boundary element such as illustrated in Fig. l-7a. An energy balance is taken on Fig. l-7a for
the first term which becomes + pwe + Basic Differential Equations (Axo “1“ fozo AZ| Kyi Ezi)h = 0 dh (Az)2 dz 2 d(pwe) dh Ax ^ dh A^\ A du (At)2 19 d(pu«) “If aA dh Az] pue + ~&rAx\ [aiAx + TzT\ * ~pue . "1 f dh At . dh "I du (At)2 A2J [d5T + ^A2JAx-pwa5 — dz | | dx 2 1 dz | 1 dx 2 — pve Ax Az = 0 where average incremental heights have been used, and hence dh dh pve = pue — h dx h = pweTz In lubrication problems, the boundaries are usually such that and thus u = w =0 I h !a pve\ =0 If the fluid has intrinsic energy at this boundary, then v ^ = 0; this may be verified by a mass balance of the same form as used above. At the zero-slope boundary pve = 0 The heat conduction term may be evaluated from Fig. 1-76 as - Kt{T - Tw) (1-22) dy _ dTdh dx dx , , dTdh + k-r- — h dz dz where the approximation, Jr.o V1 + (^) V1 + (S) - has been made. Tw is the stator-plate temperature, which may or may not be a function of x and z, and K\ is the heat transfer coefficient at the fluid-solid interface. If the runner has the same temperature distribu¬ tion as the stator plate, then dT k dy K2{T - Tw) The surface mechanical works are (UTXy + V(Ty -I- WTzy) = 0 (UTXy -{- V(Jy -j“ WTzy) 0 TXJ
20 Theory of Hydrodynamic Lubrication since u = v = w = Ow = U, and v = w =0. \h |/» \h |o ’ |o |o Substitution of the above equations into the integrated form of Eq. (l-20a) yields jl [ + hI""> ] d> - J /„' [k (‘ H) +1 (* S) ] *• -j(t££+ti£i£)l+K’<T-T-> ~ Jo ^U(Tz VTyX ^UTlZ ^ ~~ ^Txv o (l-23a) where Kt = Ki + K2. Carrying out the same operation as in Eqs. (l-20a) and (1-206) gives for the right-hand side of (l-23a): Right-hand side iide = p ju d[(u2 + v2 + w2)/2] dx + v d[(u2 + v2 + w2)/2] + ^ d[(u2 + v2 + w2)/2] | dy dz | fh (du dw\ , fh ( d [ (du dv\ -Jo p(di+^)dy-Jo |MapK^ + ^/. . d ( di/\ dp 2d (du dv dw\ L ^ 3 ^ M (to + ^ + Tz) + W Fy [" (If + %)]) dy + lo *" ^ - C/T- lo (U236) where *" - - [*(£)'■+2 (£)'-l(s + r» + S) (I + S) /dw diA2 I I ^ _L (Urn ■ ^1 ^ds: dx ,/ * ^ + dx/ dx + \dl/ + dz) dz\ + 1 And now making the approximation that since the film thickness h <$C B, L, then v « 0 and T, p j* f(y). In addition, then p and p are inde¬ pendent of y. Hence the first integral in Eq. (l-23a) may be reduced, since the continuity equation becomes d(pu) d(pw) = 0 dx dz and * = + Jc.T Also, the second integral in Eq. (l-23a) may be integrated directly. The first integral in Eq. (1-236) will cancel (for v = 0) a like integral on the left-hand side of Eq. (l-23a). With these approximations, equa¬ tion (l-23a) reduces to
Basic Differential Equations 21 + K,(T-T.)\ --, £(£ + %) i, -' f; (■ w+• ®o* - ‘u 11.+r ■" v-2w where ,, f0 /du\2 (dw\2 2 (du dw\2 (du dw\2l . OA,* * =m[2U/ +2w _n^ + */ +U+^jj (U24b) which cannot be deduced from Eq. (1-21) by stating v « 0. The next step in the reduction of Eq. (l-24a) is the neglect of any terms of insignificant size. Dealing first with terms of a mechanical nature, one of the expansion terms is by requirements of continuity, du n / p dp\ pei = 0{u7^) Hence the ratio of one of the largest viscosity terms to the expansion term is pu d2u/dy2 p du/dx W pw) which is approximately equal to 1 X 103. By taking a similar ratio with one of the terms in Eq. (1-246), uu d2u/dy2 _ n /£2\ 2u(du/dx)2 \h2) which is nearly equal to 2 X 107. It might be noted that °(M/o*uSdi/)=0(^SL) Substitution of the relations for the velocity as given by Eqs. (1-5) U = - yh) + U w = ^(y* - yh) (‘-D into Eq. (l-24a) and using the above approximations yields ([(pVh h> dp\ d(c,T) h3 dpd(c,T)~\ tL\ 2 12k dx) dx 12v dz dz J -[l(tk^) + i(hk^}} + K^T-T->)
22 Theory of Hydrodynamic Lubrication For ordinary lubricants, the characteristic values of the parameters are such that conduction is a minor mode of heat transfer. Further, the heat transfer coefficient at the fluid boundaries as reported in Ref. 5 is very small. Since it is expected that the fluid temperature rise through a bearing will not be large, the specific heat of the fluid will essentially be constant. With the above-mentioned fluid parameter characteristics, Eq. (1-25) becomes r ttt . l. T/i W dp\dT h2 dp dT~] _ \2nU> \, , /t4 (inUdx)dx (i/it/ dz dz J h ( + 12M2t/2 [(g)’ + ©II <■*> Equation (1-26) states that all of the heat generated within the fluid because of viscosity is carried away by the mass transfer of the fluid and that no heat is gained or lost through the bearing surfaces. Equation (1-26) can be made more convenient for numerical work by reducing it to nondimensional form. By setting x = x/B, z = z/Bt h = h/B, /z = m/mi, p = p/pi, V = pB/GmU, T = TpiJcvB/pJJ, where B is a representative length in the x direction, Eq. (1-26) becomes -h(i _ _ M3 dp ar = 9 m 11 , 3h4 r/ap\2 /dpV]) a dxjdx a dz dz a I P L\d*/ \di) J) (l-27a) Likewise, the Reynolds equation becomes 1-5. Equation of State. In Eq. (1-26), m is a known function of p and T and h a known function of x and z. One more equation is needed because there are three unknowns (p, p, and T), and it is provided by the equation of state given by pv = (5iT (1-28) For lubrication with a gas, the assumption that the gas obeys the perfect- gas law will be adequate. Since p and f are quantities of practical importance, the procedure might be to eliminate p from (l-27a) and (1-275) by the substitution p — 1/6(n — 1 )pT. For lubrication with a liquid film the choice of an equation of state is more difficult. Even for simple liquids the equations proposed are modified van der Waals type of considerable algebraic complexity, so that their introduction into Eqs. (1-27) would increase the complications to such an extent that the solution would probably be a major computing operation. If it be sufficiently accurate to ignore the variation of p, /z, etc., with temperature, or if the variation with temperature can be replaced by a
Basic Differential Equations 23 variation, known a priori, with x and z, then Eq. (1-276) becomes an equation for p only. The solution thus obtained can be inserted into Eq. (l-27a), which then becomes an equation in T only. This situation, which also arises in other branches of applied mechanics, enables the equations to be solved successively instead of simultaneously, but only in the order (1-276) to (l-27a). However, as soon as it becomes necessary to take variation with temperature into account, the equations become interlocked and must be solved simultaneously. SOURCES 1. Reynolds, O.: On the Theory of Lubrication and its Application to Mr. Beau¬ champ Tower’s Experiments, Phil. Trans. Roy. Soc., London, vol. 177, part 1, 1886. 2. Weilder, S. E.: Data Folder DF-54-AD-7, General Electric Company. 3. Goldstein, S.: “Modern Developments in Fluid Dynamics,” vols. I and II, Oxford University Press, New York, 1950. 4. Vogelpohl, G.: Heat Transfer in a Bearing from the Lubricant to the Gliding Surfaces, VDI-Forschungsheft, July-August, ed. B, vol. 16, no. 425, pp. 1-26, 1949. 5. Cope, W. F.: The .Hydrodynamic Theory of Film Lubrication, Proc. Roy. Soc. (London), A, vol. 197, p. 201, 1949. 6. Sternlicht, B.: Energy and Reynolds Considerations in Thrust-bearing Analysis, Proceedings of the Conference on Lubrication and Wear, I.M.E., 1957, pp. 28-38. 7. Pinkus, O.: “Counterrotating Journal Bearings,” General Electric Internal Publication R60MSD322.
CHAPTER 2 HYDRODYNAMICS OF SIMPLE CONFIGURATIONS It is not the purpose of this chapter to deal with any general problems, nor even with any specific branch of fluid flow. The subjects were chosen solely on the basis of whatever relation they may have to the study of bearings, be it as actual components of lubrication systems or as an idealization of a complex bearing geometry. Thus capillary tubes and orifices are an integral part of hydrostatic bearings, and their character¬ istics must be known before the performance of such bearings can be calculated; and the study of flow in concentric and eccentric cylinders is applicable to journal bearings and may throw some light on the flow of lubricant in the clearance space. GENERAL EQUATIONS OF MOTION FOR COMPRESSIBLE FLUIDS The three forces affecting the flow of a gas in a slot are the pressure force, the viscous force, and the force required to accelerate or decelerate the fluid. If the depth of the slot is very small and the viscous and pressure forces are very large in comparison to the inertia force, then the viscous force can be equated to the pressure force. For laminar flow this results in a simple theoretical equation relating pressure distribution, temperature, mass flow, and slot geometry. For turbulent flow use is made of the concept of resistance X, where * = \7~~T~ (2-1) Mpul, t being the “skin friction” per unit of surface area in contact with the fluid and wmvg the spatial mass velocity. The relationship between X and the Reynolds number Re, based on the hydraulic mean depth for turbulent flow, is given empirically by Blausius1-* as X - °*079 (2 2^ Xr - Ri« (2'2) and this value can be used to develop an equation corresponding to that * Such superscript figures indicate references listed under Sources at the end of the chapter. ‘24
Hydrodynamics of Simple Configurations 25 obtained theoretically for laminar flow. The assumptions made in the derivation of the flow equations are the same in both cases. The force required to accelerate or decelerate the fluid is assumed to be negligible; the pressure distribution over any cross section is assumed to be constant; and the temperature of the gas as it flows along the slot is assumed to be constant or a known function. 2-1. The Theoretical Equation for Laminar Flow. With the condi¬ tions as stated above, we can start with Eq. (l-5a). For the flow of a fluid between stationary walls we have Ui = U2 = 0, and Eq. (l-5a) becomes — hy dp u = 2/x dx giving a parabolic velocity profile. The mean velocity at any cross section is Wav 1 [h A = hJoUdy=- dx 12/x (2-3) (2-4) If the width of the slot is denoted by 6, where 6 is a function of x, the weight flow along the slot is then given by G = pgbhua, — _ &P ^ 12/i dx from which we obtain dp = _ 12 pG dx gpbh3 and by substituting in this equation the value gp = p/<HT, we obtain p dp = 12/i(R TG bh3 dx (2-5) Fig. 2-1. Elementary volume between two plates. If T, b} and h are known functions of x, this equation can be integrated and the theoretical pressure distri¬ bution so obtained. 2-2. The Empirical Equation for Turbulent Flow. Figure 2-1 shows an elementary volume of fluid between two plates, its length in the x direction being infinitesimally small and its width in the z direction being unity. The force F resisting the motion of the fluid may be considered as due to the friction acting at the surface of contact between the fluid
26 Theory of Hydrodynamic Lubrication and the wall. If the skin friction per unit area of wetted surface be denoted by r, then Fi = rAB cos 6 + rDC cos <f> that is Fi = 2r dx The pressure force acting in the direction of motion is given by ^2 = M ^ dx^j (AB sin 0 + CD sin <f>) ~(P + tx dx)(h+Txdx) = pk + (p + \txdx) Txdx - (P + txdx) (h + Txdx) which, neglecting second-order terms, reduces to F2 = -h^dx dx By equating the two forces Fi and F2, we have dp _ _ 2r dx h By substituting Eq. (2-1) we get ~ = - —(2-6) dx h The coefficient X is dependent on the Reynolds number Re. For laminar flow the value of Xl is x =24 L Re On substituting \L in Eq. (2-6), we have dp _ _ 24 uJVJ = 24 __GP_ dx Rep h Re g2pb2hz where Re = p gbp dp _ 12 pG Hence dx bhzpg , \2p(RTG , fo or pdp= bW~dx (2‘7) which is the same as Eq. (2-5). To obtain the corresponding equation for turbulent flow, we substitute in Eq. (2-6) the value \t
Hydrodynamics of Simple Configurations 27 0.079 G2 dx Re* h Re* g2Pb2h3 0.067m*G* hW'pg* , 0.067m*(RTO* , 0, pdp = Piy—(2'8) FLOW THROUGH NARROW SLOTS The general equations (2-7) and (2-8) can be integrated if b, h, and T are known as functions of x. We shall now apply these equations to various specific cases. 2-3. Isothermal Flow. Constant Area Slot. For a constant area we have h = const and b = const, and integration of (2-7) gives for steady laminar flow 2 2 24n&TG . . fo m Pi2 - gp— (*2 - *i) (2-9) On integrating Eq. (2-8), we obtain for turbulent flow 0.133„*(RTO»(*. - X,) P> “ P2 = pwp ( ) Diverging Width. Let us denote the rate of width divergence by b — ax. Then if the frictional effect of the two side walls is neglected, the flow is analogous to the radial flow between two circular flat plates of constant film thickness, and by symmetry the pressure at any given radius is constant. Provided, therefore, that the divergence is small, the pressure in any plane perpendicular to the axis is approximately constant, and Eqs. (2-7) and (2-8) may be applied. By substituting in these equations b = ax, we obtain for laminar flow , 12 p(RTG dx pdp = u ahz x which when integrated becomes 2 2 24M(Rrci x2 /011N (2'n) and for turbulent flow , 0.067m*<R7’G* dx pdp ^ which when integrated becomes \Xi« Xi’V . 0.178x xi P* P2 ct/'h3gii I •*••* ) (2-12)
28 Theory of Hydrodynamic Lubrication Diverging Depth. Let us here denote by h = fix (b = const) the divergence of slot depth. Because the divergence of the two plane sur¬ faces is very small, the component of velocity perpendicular to the axis of the slot must also be very small, and the pressure distribution over any cross section is, therefore, assumed to be constant. For laminar flow, substitution of h = fix in Eq. (2-7) yields , 12 dx Vdp and, when integrated, this becomes Pi2 - p2 W3 \Xi2 X22 By the same substitution in Eq. (2-8), we have for turbulent flow 0M7n*(RTGK dx (2-13) p dp = — which when integrated gives blip SgK t 0.067MM<nrow / 1 1\ /01,, Pl -w ~ &) ( } 2-4. Flow through Orifices in Series. Let us consider two identical i ft, Pbi | | T Po^ai | 1 \h 1 ..I Lj Ps 1 Pb o b (<7) \b\ Fig. 2-2. Orifices in series. orifices a and b as shown in Fig. 2-2o. The pressure ratio across a and b is „ _ Pal ^ _ Pb2 ra Tb Pal Pb1 where pb i = pa 2 The flow for a single orifice based on the perfect-gas equation is given by G = CA [2g (^j (r2/* - j* (2-15) where vt — specific volume at supply condition k = isentropic expansion coefficient From continuity of weight flow we have Ga = Gb. Thus, by assuming that the areas and discharge coefficients are equal and expressing the
Hydrodynamics of Simple Configurations weight flow by Eq. (2-15), we have TaW{ 1 - ris/*(i _ rt<i-*)/*) Val Va2 (2-16) 29 If isothermal flow is assumed, we get Vg2 _ Pal Va\ Pa2 On rearranging the terms and substituting the above equation into Eq. (2-16), we have By knowing the pressure ratio across one orifice, the pressure ratio across the other orifice can be found from Eq. (2-17). It can be easily shown (Fig. 2-26) that, for small area ratios, the velocity at the last orifice is given by If the area and discharge coefficient of each orifice are given by A and C, respectively, then the volumetric flow through the last orifice downstream is given by If the pressure ratio across the last orifice is equal to or less than the critical • (0.53 approximately), then the volumetric flow is a constant and is given by The flow through the next orifice upstream (which is not choked) is given by Eq. (2-15) and is or Qi = CAUi Qi = CA lig ptf! T1 - Jh r = 2? < 0.53 Gi = gpiQi = gpiCA 2g - (0.53)<*-»'*]|H (2-18)
30 Theory of Hydrodynamic Lubrication If we equate the weight flows through the orifices and still assume that the areas and discharge coefficients are equal, then [ 1 - (0.53)(t-l)/‘l = ^ (r22'* - r2<*+>«‘) Vl L J v2 Assuming again that pv = constant, the above equation reduces to r2_2(r22/* - r2«+l)lk) = 1 - (O.W'~»lk (2-20) From Eq. (2-20) we see that the pressure ratio r2, and therefore any pressure ratios upstream, will be constant. Therefore, the flow from the inlet up to the orifice with the pressure ratio r2 will be choked. Under Table 2-1 Number of orifices in series Approximate critical pressure ratio Critical supply pressure (exhaust to atmosphere), psia 1 0.528 27.8 2 0.430 34.2 3 0.370 39.8 4 0.330 44.5 5 0.305 48.2 6 0.290 50.7 7 0.275 53.5 8 0.265 55.5 9 0.255 57.7 10 0.250 58.8 these conditions the pressure upstream of the last orifice is directly pro¬ portional to the supply pressure. Looking at Eq. (2-18) we see that the weight flow through the last orifice is equal to gpi(piVi)** times a constant factor or Gi — gpi(piVi)^K K - const Now, since gpi = \/v\, the above equation reduces to «■ - {v)“K From the perfect-gas law we have = pi Thus we get a - ”• (irrl)"x
Hydrodynamics of Simple Configurations 31 Since pv = const, GiT = const and we have G\ = piKi K i = const We have already deduced that pi/p, = const; therefore, the equation above becomes G\ = p,K2 — < critical volume K2 = const P* and we see that the weight flow through a series of orifices with the exhaust orifice flow choked is directly proportional to the supply pressure, provided that the discharge pressure is constant. For a discharge to the atmosphere, the critical pressure ratio and critical pressures are given in Table 2-1. INCOMPRESSIBLE FLOW 2-6. Flow between Parallel Walls. The simple case of laminar flow of an incompressible fluid between two parallel surfaces of infinite extent is given by Eq. (l-5a). By placing the coordinate axis y — 0 at the center of the film, i.e., at h/2, and setting U2 = 0, we obtain -?(• + *!)-si [>-*©'] For the case U = 0, that is, when both surfaces are at rest, Eq. (2-21) gives a parabolic velocity profile; for dp/dx = 0 and U ^ 0 the velocity profile is linear. For the general case the velocity profile is made up of both contributions, and, depending on the value of dp/dx, the profile may be convex or concave, or it may even reverse itself. The point at which flow will occur in a direction opposite to U can be calculated by letting du/dy = 0 at the stationary wall. The result is that, for dp/dx > 2 Up/h2, the flow will reverse itself over certain values of y. The general behavior of the velocity profile is shown schematically in Fig. 2-3. If the two surfaces —h/2 and +h/2 are kept at uniform temperatures of T2 and 7\, respectively, then, since dT/dx = 0, we have from the energy equation . d2T (du\2 dy1 M \dy) By using du/dy from Eq. (2-21), d*T _ IPp [ 16umtkXy (twyVl 'dy2 h2k L Uh ^ \ Uh ) J where umAX = — ^ is the maximum velocity. 8p dx
32 Theory of Hydrodynamic Lubrication The solution of the above equation is = t, + (p- - T 2 . fiUUn 3 k 1 + 2 fiUum 3 k The temperature gradient is dT = Ti - T2 , 2mUum dy h 3kh [i+80D1+tHi _ i6©' IS 2hk I \h) (2-22) + 16 um ~U \h) , 128wL 3 U2 \h (2-23) and depending upon whether the first right-hand term of (2-23) is greater or smaller than the remaining two terms, heat will either flow into or out of the upper wall. T->0 dx du I dy I > o ± = o dp dx dx dx du = 0 1 h £1 >° dy | h du dy dy | 1 * 2 2 2 < 0 > 0 2-6. Circumferential Flow between Concentric Cylinders. The lam¬ inar flow between concentric cylinders extending to infinity in the axial direction yields from physical considerations u '= u(r) v = w — 0 p = p(r) When these considerations are used in the Navier-Stokes equations, the following two differential equations are obtained dr2 ^ dr \r) pu2 _ dp r dr (2-24 a) (2-24 b)
Hydrodynamics of Simple Configurations 33 By using the boundary conditions of Fig. 2-4, we obtain by integration of Eq. (2-24a) U = [r(u>2ft22 - Ulft,2) - («2 - «,)] (2-25) and for the pressure distribution from Eqs. (2-25) and (2-246) * = PI + (g,»-B1y[W - — 2i?l2/?22(&>2 — <*>i) (o^fi^2 — Wj/?12) In til 2. where pi is the pressure at Ri. -W«i« fi2‘(«. - »,)* ^ f i)] (2-: 26) Fig. 2-4. Notation for concentric cylinders. If the inner cylinder is kept at rest, the moment of the fluid on a length L of the outer cylinder is M 2 = 2ttR22Lt2 — 2tR'?Lp,t «GD or Af, = 4irL “2 (2-27) This last equation can be used to determine m by merely measuring the moment M2. When only one cylinder rotates in an infinite fluid, we have for o>2 = 0, T — oo /?l200l w = r Af = ^tcpLR\2<j)i The energy equation for our case is H / dT\ = _ m /dw _ w\* r dr \ dr / A; \dr r /
34 Theory of Hydrodynamic Lubrication which gives the following temperature distribution T—T I M ^i4^24(o>i — 2)2/ 1 1\ 1_1"/b (RJ-RfY \RS r2/ T T fl Rl4R24(d)l-U)2)2 ( 1 l\ l2^k (RS-RW \RS RS), r In Rt/Rx Ri (2“28) 2-7. Axial Flow in Cylinders. Concentric Cylinders. The differential equation for the axial flow of fluid in an annular space is + <2'29> By using the boundary conditions of w = 0 at Ri and R2i where Ri and R2 are the inner and outer radii, the velocity profile in an annular slot becomes + (2-M> By integrating w between Ri and R2, the flow becomes «-*■•>[*1’-ft’-ETIGTO] <M1> By setting Ri = 0 in Eq. (2-30), which gives the flow in annular slot, we obtain the flow of a liquid in a circular cylinder. The velocity profile is given by (2-32> 1 2 w-“"i»TzR> The velocity profile is parabolical, therefore 1 1 dp p 2 W.v, - 2 W,„ gM dz Ri The volume flow is given by Q = RS - ^ (2-33) The energy equation is given by . /d2T , 1 dT\ /dw\2 \dr2 r dr) M \dr) so that for a temperature of T2 at R2 and at r = 0 we have T-Tt+ (Ti _ T,) [ 1 - (2-34)
Hydrodynamics of Simple Configurations 35 Equation (2-33) is useful in the calculation of the performance of inter¬ nally pressurized bearings where lubricant is admitted through a series of capillary tubes. Flow through Eccentric Cylinders. If the cylinders are not concen¬ tric, the slot height h is a function of 6 and is given by the equation C(1 + c cos 6) (Chap. 3). If we are allowed to simulate the case of eccentric cylinders by two nonparallel developed surfaces, then the velocity from Eq. (1-56) is given by and now is a function of both coordinates. The flow is, by integrating w, For c = 0, the above equation reduces to (1-56). The exact analytical treatment of flow in eccentric cylinders is compli¬ cated. It represents a two-dimensional problem which is described by Poisson’s equation where w — 0 at the boundaries. A partial solution of this problem is given in Ref. 5, where the velocity profile is expressed in series form by and the various terms are, referring to Fig. 2-5, defined as follows: C = distance from the origin (z = 0) to either pole of a bipolar coordi¬ nate system w = - ^ y[y - C(1 + € cos 0)] (2-35) Q = (1 + ««*) (2-36) d2w dhv _ I dp dx2 dy2 p dz (2-37) where C + z p = q~zt~z 2 a variable in the complex plane
36 Theory of Hydrodynamic Lubrication An interesting result reached by evaluation of Eq. (2-37) is that the velocity distribution at any section in the variable-height annulus is very close to the one that would have resulted from a concentric case with the clearance equal to the dimension of the particular section under consideration. Equation (2-35) is thus a good approximation of the actual velocity profile. SOURCES 1. Piercy, N. A. V.: Aerodynamics, Elektrotech. u. Physik., p. 278, 1937. 2. Shires, G. L.: The Viscid Flow of Air in a Narrow Slot, ARC Tech. Rept., Cp 13 (12329). 3. Robinson, G. S. L.: Flow of a Compressible Fluid through a Series of Identical Orifices, ASME Paper 48-APM-4. 4. Shih-I-Pai: “Viscous Flow Theory,” vol. I, D. Van Nostrand Company, Inc., Princeton, N.J., 1956. 5. Poritsky, H., and Fend, F. A.: Laminar Incompressible Flow between Non-con- centric Circular Cylinders, TIS Repl., 57GL54, General Electric Company.
CHAPTER 3 INCOMPRESSIBLE LUBRICATION; ONE-DIMENSIONAL BEARINGS With the exception of Chap. 6 we shall be dealing throughout this book with hydrodynamic lubrication. By “hydrodynamic lubrication” we mean a process in which two surfaces, moving at some relative velocity with respect to each other, are separated by a fluid film in which forces are generated by virtue of that relative motion only. As in all other problems in engineering, the solutions on the following pages are based on certain assumptions, and in order to appreciate the degree of applica¬ bility of these results, a realistic picture of bearing operation will first be given. THE REAL BEARING Figure 3-1 shows a journal bearing operating with an external load W and speed (7. Under the physical conditions imposed, the journal will run at some eccentricity e, the region below the line of centers 00' form¬ ing a converging and the region above 00' a diverging space. From A to D the lubricant is being pumped by the journal into an ever-decreasing space with the result of building up high pressures in the fluid. The eccentricity e and the magnitude and distribution of these pressures will be such as to yield a resultant force equal and vectorially opposite to W. In the process the fluid is being continuously squeezed out the ends of the bearing, and this side leakage, plus any conduction and radiation that may exist, carries away the heat generated by the rotating journal. New lubricant is being delivered at point B to replenish the amount lost by side leakage. The hydrodynamics of thrust bearings are essentially represented by Fig. 3-2. Available bearing solutions even in their elementary form satisfy the basic requirements of continuity and momentum and express the per¬ formance of bearings as a function of load, speed, viscosity, and bearing dimensions. However, there are other significant features which are often disregarded either because of incomplete knowledge or because of mathe¬ matical difficulties. Together with the basic theory they constitute the physical reality of journal and thrust bearings. These features are: 37
38 Theory of Hydrodynamic Lubrication 1. Boundary Conditions. The pressure profile in journal bearings starts at the point where lubricant is admitted if 0i > 0 and at 0 = 0 is 0i < 0 ( —ir < 0 < ir). In a full bearing the pressure profile ends beyond hmiD, at 02, where it falls to a value very slightly below that pre¬ vailing at the bearing sides and then rises again to equal the boundary pressures. In those partial bearings where the arc ends before E and in most thrust bearings the pressure profile ends at the exit, much as shown in Fig. 3-2. The dip at the end of the pressure wave, region EG} Fig. 3-1. Dynamics of a full journal bearing. cannot be eliminated by simply raising the inlet pressure at B. Com¬ pared with the pressures prevailing from B to E, the values in region EG are negligible. 2. Striation. The fluid film in full journal bearings is rarely complete. If a deep axial groove is cut at B, the full film will start along B. If there is only a hole for admitting the lubricant, a full film will form along the dashed line B' of Fig. 3-Id. Between B and C the flow, because of the unfavorable pressure gradients, will consist of the shear flow less the pres¬ sure flow. At C, dp/dd = 0, and only shear flow prevails. From C to E, the flow consists of shear flow plus the pressure flow. Past hmin the clearance space begins to increase. The extra flow available from the pressure component at D will help fill out the increasing space until the flow is equal to the shear flow at F. From F on, the clearance continues to increase, and since there is not sufficient fluid to fill it, the film breaks
Incompressible Lubrication; One-dimensional Bearings 39 down into individual filmlets and continues in that state until fresh lubricant is admitted. The space between the lubricant filmlets is filled with air, vapor, and foam. This is schematically shown in Fig. 3-1 d; Chap. 15 offers experimental evidence of this phenomenon. In partial bearings the situation is similar if the arc ends beyond F. If the trailing edge is before F, the film is complete throughout the bearing. The films are usually complete in thrust bearings. 3. Viscosity. The viscosity of the lubricant in hydrodynamic bearings never remains constant. The viscosity of any fluid varies with both temperature and pressure, and there is also evidence that it varies with the rate of shear. While the variation with pressure is significant only at very high pressures, usually beyond ordinary bearing operation, the u~-U Fig. 3-2. Dynamics of a thrust bearing. dependence on temperature is most pronounced at low and moderate temperatures, the very regions in which bearings operate. When losses are low or temperature levels high, average values may be used. When the conditions listed above are at the other extreme, constant viscosity values may yield unsatisfactory results. 4. Heat Transfer. Not all the energy generated in a bearing is carried away as heat by the lubricant. Part of that energy is dissipated by conduction and radiation via the bearing shell, housing, and journal. No two bearing assemblies are alike in this respect; temperature variation over each of the mating surfaces, the presence of neighboring heat sources and sinks, the complexity of assembly parts, and the effect of windage create a formidable problem, particularly since the hydro- dynamic and the heat transfer problems are interrelated and have to be
40 Theory of Hydrodynamic Lubrication treated simultaneously. Usually, the smaller the bearing and the lower the shear losses the higher the percentage heat lost to the surroundings. 5. End Effects. This subject is still unexplored, but a multitude of experimental data on lubricant flow which refuse to conform to theoretical predictions can be explained only by the effects to be mentioned. These include phenomena such as surface tension at the sides of the bearing, the formation of a meniscus, and the sealing effect that such a meniscus has on the free flow of lubricant out of a bearing. This sealing mecha¬ nism causes the lubricant to flow backward along the sides of the bearing, i.e., in a direction opposite to journal rotation, and to reenter the bearing in the low-pressure region to be recirculated. Another phenomenon is the possible formation of a vena contracta around the annular outlet. These and perhaps other effects make the side leakage usually less than that predicted from theory. The points listed above are all major and general phenomena associ¬ ated with hydrodynamic bearings using incompressible fluids. Some points that assume significance only in certain ranges of operation are: a. Elastic Deformation. Under heavy loading and depending on its structure and assembly, the bearing surface will deform. This in effect will produce a different film shape with a drastic change in bearing performance. b. Turbulence. The Reynolds equation is based on the assumption of laminar flow. High linear speeds, large clearances, and low viscosities will cause turbulence with a resulting rise in power loss, a drop in lubri¬ cant flow, and a shift in the locus of shaft center. c. Thermal Expansion. When bearings undergo appreciable tempera¬ ture changes, when journal and bearing materials have radically different coefficients of thermal expansion, or when journal and bearing are forced to expand against each other, the clearances will not retain their original shape and dimension, and the performance of the bearing will be affected. d. Surface Roughness. When bearings are operated at very low values of hmin, the inherent surface roughness of all materials may have an effect, since the roughness may be of the same order of magnitude as the mini¬ mum clearance. This may not only change the shape of the oil film but carry the operation into a mixed boundary region where the require¬ ments of hydrodynamic lubrication are no more than partially fulfilled. e. Unbalance. Most journals will have some residual unbalance. With unbalance the journal center is not confined to a point but moves along some locus. Thus, the steady state is replaced by dynamic conditions. /. Misalignment. Slight amounts of misalignment are inherent in all journal bearing assemblies. When the degree of misalignment becomes excessive, it is necessary to take this effect into consideration.
Incompressible Lubrication; One-dimensional Bearings 41 Very minor items which do not affect bearing operation to any notice¬ able degree but of which one should be aware are the variation of specific heat with temperature and pressure and the presence of air, foam, and foreign particles in the lubricant. In the solutions of this chapter, none of the points mentioned above is considered. In fact, all these ramifications are minor compared with the radical and from a practical standpoint impossible assumption of a one¬ dimensional bearing, a bearing infinitely long or infinitely short. These solutions, however, are useful for a number of reasons. In the first place, they are given mostly in analytical form with all the inherent advantage over numerical results, of which the bulk of exact solutions consists. Oftentimes they are the only available solutions, a useful guide to how a bearing would possibly perform under certain conditions, and they do provide upper or lower limits. Although the quantitative answers are often at variance with experimental results, they do nevertheless provide a means of studying trends and relationships. ONE-DIMENSIONAL JOURNAL BEARINGS If we assume the bearing to be infinitely long in the axial direction, this implies no variation of pressure in the z direction, or dp/dz = 0. Equa¬ tion (1-12) then becomes (3-1) dx \p dx) dx If the flow due to the pressure gradients in the x direction can be neglected (while retaining the com¬ ponent due to shear)—a situation approached by very narrow bear¬ ings—then the term £C-‘2)-° and Eq. (1-12) becomes dz\p dz ) dx (3-2) It should be noted that this last equation is still a function of two vari¬ ables and that it is a less radical simplification than Eq. (3-1). The above equations are fully defined except for h. While the film thickness in thrust bearings can assume different expressions, its form for an aligned journal bearing is universal. Referring to Fig. 3-3, we have
42 Theory of Hydrodynamic Lubrication OB _ R _ R _ e sin 0 sin (7r — 0) sin 0 sin a P = 0- a = 6- sin-1 sin 0^ OS = ^ sin |^0 — sin-1 ^ sin 0^j = R2 — e2 sin2 0 — e cos 0 h = (S + C) - OB = C + e cos 0 + R — \/#2 — sin2 0 « C + e cos 0 or h = C(1 «+• e cos 0) (3-3) 3-1. Infinitely Long Bearing. The earliest solution of the infinitely long full journal bearing is due to Sommerfeld, who by use of an adroit substitution succeeded in integrating Eq. (3-1). When x is replaced by the angular coordinate 0, that is, when x = Rd, and it is remembered that n is constant, Eq. (3-1) becomes WR% (3-4) de\ de) By integrating once with respect to 0 dp de ~ 6 nURh + Ci h3 II ^3 1^3 0, so Ci II 1^3 6 nURh — ho h3 At some h = ho, Ann h 2i - (3-5) where h is given by Eq. (3-3) and ho is still to be determined. From Eq. (3-5) the pressure is given by _ 6nUR \ f dd h0 f d$ 1 r P C2 Li (1 + e cos 6)1 C J (1 + « cos 0)3J + 2 To integrate the above, let 1 *> 11 n 1 — 1 -f € cos 0 = , 1 — c cos ^ ... . COS \p — € from which cos 0 — 1 — e cos By using sin2 0 + cos2 0 = 1, we have (1 — t2)Yi sin \j/ sin 0 = 1 — c cos \J/ and by differentiating one of the terms above (l-€2)^#
Incompressible Lubrication; One-dimensional Bearings 43 The boundaries 0 = 0 and 0 = transform into the same boundaries in the ^ coordinate, and thus the boundary conditions are p — pa at \f/ = 0 p( 0) = p(2t) (3-6) By evaluating the integrals resulting from the above substitutions, we have / de (i +1 cos ey _ r n -1 cos a» a -€%« J V 1 - €2 J 1 — 6 CO" *#‘ and Thus / d0 € COS ^ d\p = (1 _V)W (* - 2t sin^ + e-J +J8in2^ _ /* /1 — € COS A2 * (1 + € COS 0)2 “ J V l - €2 / (1 - w* ho = n~ 4 sin ®(a) = C2 [ (1 - «2)* C(1 h 0 _ - «2)« V 2e sin \p c2 sin 2^ j 4- C 2 By using the first boundary condition of Eq. (3-6), we have C2= pa By using the second, we get 2C(1 - €2) ho — 2 + €2 (3-7) and thus the expression for the pres¬ sure distribution becomes, by revert¬ ing to the original coordinate, P = Pa + 6pURe (2-f« cos 0) sin 0 C2 (2 + €2)(l + 6COS0)2 (3-8) Fig. 3-4. Pressure distribution in¬ cluding negative regions. where pa is the pressure at 0 = 0. This can be evaluated from the condi¬ tions at B, where a given inlet pressure pi corresponds to a given angle 0i. If the inlet hole is at 0i = 0, then of course pa is the value of inlet pressure. Equation (3-8) yields regions of high negative pressures such as shown in Fig. 3-4. The magnitude of these negative pressures will depend both on the position of 0i and on the magnitude of p\. In any case, the
44 Theory of Hydrodynamic Lubrication pressure distribution resulting from Eq. (3-8) is always antisymmetrical about 0 — t and p = pa. - "• { The vertical load component is/by integrating the pressure over the bearing surface given by W sin <*. = jf" LR de p sin 6 By integration by parts, W sin <f> = LR £ — p cos 6 — j ^ cos 0 dd j and, by using from Eq. (3-5), we have dd cos 6^d$ du . J 6^LUR2[h0 f2' cos 9 de f2 wsm* = —c^[c]0 (i + (cosey ~ Jo 2r cos 6 dd (1 + C cos 0)2J These can be reduced to the same integrals used in evaluating the pressure distribution by writing cos Q _ 1/c 1 /€ (1 4- c cos 0)z (1 + € cos 0)2 (1 + c cos 0)3 , _ cos 0 _ 1/c 1/c an (1 4~ c cos 0)2 (1 + « cos 0)2 1 4" c cos 0 Thus -«») f2r ^ C2 \ 24- €2 > (1 4- € cos 0)3 d0 : COS 0 | w ^ 6nULR* (2(1 - c2) f2* dd W sm <f> = — j v ' 1 _ f 2(1 — t2) ] [* de [» dfl L 2 + «* J Jo (i + * cos ey Jo i +1< The only new integral to appear is the last one, and its value is de _ i / 1 4- € cos 0 y/\ — e2 By following the same procedure used for evaluating the pressures, we obtain 1V sin 6 - (3 „x w sin <t> ~ (2-|- e2)(i _ t2)W ^ To find the attitude angle, we must look for the load component at right angles to W sin <t>, or W cos <j> = j^'LRde p cos 0
Incompressible Lubrication; One-dimensional Bearings 45 Upon integration by parts, W cos <t> = LR sin 0 — j sin 6 ~ dd j = — LR sin 6 ^ dd By use once again of the Sommerfeld substitutions, ^ GpULR*\ ho ( # c2 . , A . 1 .1 I2' W COS <t> = Ci i) (cos j, - 2- Sill2 + J— COS |o and W cos <f> = 0 or, since W ^ 0, <f> = ^ (3-10) and the displacement of the shaft is always at right angles to W sin <f>. Moreover, since there is no load component at right angles to W sin 4>, W sin <f> = W is the total resultant load. This unrealistic result is a consequence of including the negative pressures in the integration for load capacity. Thus we can rewrite Eq. (3-9) as _pN/R\_ p Vcj " (2 + e2)(l - t2)» . 12rt W'11) and the Sommerfeld number is seen to be a function of c only. The shear stress at the journal is, from Eq. (l-16a), given by — . h dp Tx ~ h + 2R d§ and the frictional force on the journal is de Fj = j*’r,LR de = -LR J*' j- + € COS d M /02'( 1 + ec°s e^de\ By using for the integrals the expressions derived above F - i tt # 4x(l + 2c2) Fj pLL c (2 + €2)(1 _ e2jW The friction factor defined as / = F/W is then f-h-9. I + 2i! (3-13) W ~ R 3f W) At the bearing surface, by Eq. (l-15a),
46 Theory of Hydrodynamic Lubrication The difference in the journal and bearing torques is balanced by the external load IT, which exerts a moment through its eccentricity e or RFj = RFb + We (3-14) The friction in a concentric journal bearing when e = dp/dO = 0 is often referred to as Petroff’s equation and is given simply from Eq. (3-12) by 2 thULR/C, The foregoing analysis yields quantitative results that are far afield from any actual bearing performance. They are also qualitatively in error; for, as mentioned previously, negative pressures of the same order as the positive pressures could not possibly be maintained and the locus of shaft center is never a straight horizon¬ tal line. The major shortcomings of the foregoing analysis can be eliminated by imposing a more realistic boundary condition at the trailing end of the pressure wave. These boundary conditions are discussed in greater detail in Chap. 4. Suffice it to say here that no negative pressures are allowed and the requirement is imposed that, at the point where the pres¬ sure wave falls to zero, line E, the pressure gradient too becomes zero, as shown in Fig. 3-5. Thus the boundary conditions imposed are: vq CTmin Fig. 3-5. Pressure distribution ex eluding negative regions. p = 0 dp _ dd ~ p = 0 = 0 at 6 = 0 at d = d2 at 6 — d2 (3-15) Since Eq. (3-1) is a second-order differential equation, it cannot in general satisfy more than two boundary conditions. It will be shown, however, that the last two conditions of Eq. (3-15) are a special case of a more general single condition and that the solution based on Eq. (3-15) is only one of a family of possible solutions. Let the last two boundary conditions at 02 be written as the single condition D = P kd9 then, by integrating the Reynolds equation, we obtain QnUR[ f9 dd , „ f9 dd • V = Writing C2 —-7^—3T2 + Ci je> , r.i- + Ci [ (1 + 4 cos 9) -/: * (1 + € cos 6)n (i + € cos ey i_ 1 + € COS 02
Incompressible Lubrication; One-dimensional Bearings 47 the constants C1 and C2 are evaluated by using p = 0 at 0i and P = fcgf atfl2 n h + kg2 n _uglh-9ih tl ~ h + kg* °2 k h + kg* The expressions for p and dp/de then become 6mI/K P C2 f* de _ It + kg* [• Je, (1 + < cos BY h + kga Je, de (1 + € cos 6 9*1. dp 1 /2 "h kg d0 C2 [(* + € cos 0)2 ^3 + kg3 (1 + e cos 0)3J Since e < 1, 7n > 0, and g > 0, the denominators never vanish. Thus the conditions for p = 0 and dp/dd = 0 at 0 = 02 are for p = 0 k(g2h - gzI2) =0 or %2(/3 - ^/2) = 0 for5? = 0 0(/2 + kg2) = /3 + kg3 or /3 - fif/2 It follows then that, to have p = 0, we need either k = 0 d0 or /**« d0 /'®I / /i _i_ —zr2 = (1 + * cos 02) / Jet (1 + € cos 0)2 J9l (1 + c cos 0)3 Thus if the later condition holds, both p and dp/dd are zero for any value of k including k = 0, which corresponds to our particular boundary conditions. By again employing Sommerfeld’s substitutions and using the first two boundary conditions of Eq. (3-15), we obtain for the pressure distribution 6 nUR I, . (2 + «2)^ — 4e sin \J/ -J- e2 sin $ cos yp ] P = C*( 1 - «»)* T " * * 2[1 + e COS {h - r)\ ) (3-16) where cos ^ = € + cos 0^ an^ ^ corresponds to 02. By using the last 1 *i € COS tt condition, namely, p = 0 at ^ = ^2, we obtain from Eq. (3-16) €[sin (^2 — t) cos — x) — ^2] + 2[^2 cos (^2 — x) - sin (*2 - tt)] = 0 (3-17) which determines ^2 and thus 02. Equation (3-16) with ^2 determined by Eq. (3-17) gives a pressure profile satisfying all the conditions of Eq. (3-15).
48 Theory of Hydrodynamic Lubrication For the two load components, by writing ^ = ^2 — a.,,, (1 - «2)(1 + t COS^'j) W sin <b = WW/WW* cos # ~ sin #) . . ►V sin </> (1 _ + { cog (6 19) 3nUL(R/cr r ««(i + cos (1 - «2)*(1 + t cos f'2) [ 1 - e2 + 4(^2 cos i/'i — sin ^2)21 (3-20) 2(1 — <2)^(sin ^2 — *p2 cos ^2) «(1 + cos i.„„ ^ c r vam ^2 Y2 WO Y2J /o m\ tan <f> — .//\2 (3_21) («)/=^± + _^!s_ (322) The use of conditions (3-15) resulted in the elimination of the region of negative pressures and the derivation of a journal locus, Eq. (3-21), which conforms with experimental evidence. Numerical results of Eqs. (3-20) to (3-^2) are given in Table 4-1, where they are tabulated together with the solutions of finite bearings. 3-2. Infinitely Short Bearing. Equation (3-2) has been written down as applying to infinitely narrow bearings. Since for aligned journals h = f(x) alone, we can integrate this equation by treating h as a constant. Thus QnUdh z2 h3 dx 2 _ uflu an z- r r V = TT + C'z + C* By using the boundary conditions p = 0 at +L/2 **•■> - W- (t - *■) ir-r^'w <3-23> This pressure distribution is parabolic in z and antisymmetrical about 6 — nr and p = 0; that is, the region of negative pressures is identical to that of positive pressures. Here the problem of negative pressures is dealt with simply by deleting the region t < 0 < 2tt where the negative pressures occur. (This, of course, can also be done for the infinitely long bearing, and results of such an integration are given in Chaps. 7 and 12.) By summing forces only over the interval 0 < 0 < tt: c sin 0 cos 0 , ad Wx = —2 f f p cos 0 R d0 dz = — [ n , jo Jo 2c2 jo (i +«cos ey o [' fL/2 a T> Ja j nUL* [* € sin2 0 Wy = 2 \ / p sin 0 R dd dz / y—- -r- d0 Jo Jo 2c2 jo (i + € cos ey By again using the Sommerfeld substitution and integrating,
Incompressible Lubrication; One-dimensional Bearings 49 (3'24a) W‘ = !^W^W> (3-246) The total load capacity is then given by pUL3 e 4C2 (1 - €2) ^ ,t - .M I'**1 - ‘2) + 16‘2]» (3-24c) T (§)’ ’ s (I)’ - .■)' + I6.'jw <«-25> This last expression is seen to be independent of bearing diameter. The effect of diameter is felt through the value of Cy which is usually a function of bearing diameter. The attitude angle is given by tan 0 = ^ ^ ~ (3-26) Since there is no pressure-induced shear, _pU T h and F = J*' M £ LR de = (3-27) with the friction coefficient given by — f — 2t2£ (o og\ CJ “ (1 - €2)^ ( } The lubricant flow out the sides of the bearing is, by Eq. (1-136), f'RVdp Qz~ Jo 12m dz d6 = eULC (3-29) ±L/2 The parameters at the point of maximum pressure are 1 - (1 + 24e2)>* COS do = 4e ho = j [5 - (1 + 24*2)»] (3-30) _ _3pUL2 c sin 0O /b oi\ P° 4RC* (1 + t cos fio)3 1 ' This treatment yields a fair approximation to the performance of narrow bearings at low eccentricities and, by the simplicity and compactness of its mathematics, constitutes a useful tool in the analysis of lubrication problems.
50 Theory of Hydrodynamic Lubrication 3-3. Partial Bearings. By definition any bearing having an arc less than 2t is a partial bearing; in practice, however, the criterion is usually an arc less than 180°. The analysis of partial bearings, as evident from Fig. 3-6, is made more difficult by the appearance of two new independent parameters, the load angle a and the arc span 0. Most available solu¬ tions are similar to the Sommerfeld solutions for a full bearing and in¬ clude negative pressures. We shall, however, restrict ourselves to bound¬ ary conditions similar to those of Eq. (3-15), namely: V = Pi dp n v = Te =0 at 0 = 0i at 0 = 02 p > 0 in the region 0i < 0 < 02 p = 0 at 0 > 02 Although the methods and the equa¬ tions involved are similar to those of full bearings, the actual calculations are laborious and the final expres¬ sions are long and cumbersome. For this reason, the results will be given in graphical form. If in Eq. (3-5) the limits of integration are kept general, then P = Pi + 6/tiUR feh- h0 Je, hz or after performing the integrations in the same manner as before V = Pi + rf‘ 6 pUR C2(l -62)* {[(1 <2) c(1 + 2)](* ^ - £e(l - «2) - ^T-°j (sin - sin f,) - (sin 2^ - sin 2^i) (3-32) where \p is the angle from Sommerfeld’s substitution and h0 is to be evaluated from the boundary conditions at 02. Equation (3-32) was evaluated by numerical integration for low eccentricities and by the use of the mathematical expression for high eccentricities. The results are extracts from families of solutions and represent points of maximum load capacity, a condition at which all bearings tend to operate. Figures 3-7 and 3-8 present performance of partial bearings with the load vector located at any arbitrary position with respect to the bearing boundaries
Incompressible Lubrication; One-dimensional Bearings 51 and containing no negative pressures. For a given S and arc length 0 Fig. 3-7 will, if 0 is set equal to 0', determine the optimum load angle a. If 0 > 0', then obviously the load ca- Numbers along curves indicote attitude angle <f> 150° 180° >3 = 210° pacity will be the same, with the arc 0 — 0' only increasing the frictional losses. The condition 0 = 2a rep¬ resents, of course, the case of cen¬ trally loaded bearings. It should 22 r 0 20 40 60 80 100 120 140160 180 200 Load angle a, deg Fig. 3-7. Load capacity of partial bear¬ ings. 0 20 40 60 80100 120140160180 200 Load angle a, deg Fig. 3-8. Friction factor in partial bear¬ ings. be pointed out that here, as throughout the book, the reference area for partial bearings is the projected area of an equivalent full bearing, that is, P = W/LD. 3-4. Fitted Bearings. Fitted or no- clearance bearings are those whose diameter equals the diameter of the journal, as shown in Fig. 3-9. In that case the clearance is zero and Eq. (3-3) becomes h = e cos 0 Fig. 3-9. Notation for fitted bearings. By rotating the 0 axis by 90° in the di¬ rection of rotation, using h = e sin 0 in Eq. (3-4), and expressing the constant ho by e sin 0O, we obtain p(9) = j^sin 0o - In tan - 2 cot 0 j + <7, By using the boundary conditions p(0i) = p(0o) = 0, we obtain for (3-33)
52 Theory of Hydrodynamic Lubrication sin 0O and Ci 2(cot fli — cot 02) Q 1/1 1/ V2J /Q Sin 0 cot 0i/sin 0i — cot 02/sin 02 + In [tan (02/2)/tan (0i/2)] n 3nU R [8in 00 (s§4l “ln tan I1) "2 cot *i] The integrations for load capacity performed over the entire bearing arc yield the following results: SnULR2 T -Cie2 , . i • n ( n i , 02 Wy — ^ SjHJR ^C0S — cos 2' sm 0 I cos 2 2 — cos 0i ln tan —^ — 2(sin 02 — sin 0i) j (3-35a) Wx = (s^n ^2 sin ^0 + sin 0O ^cot 0i — cot 02 0 0 \ + sin 0i ln tan ^ — sin 02 ln tan + 2 ^cos 0i — cos 02 + ln tan ~ — ln tan ^ j (3-356) j^4 ^ln tan |-2 — ln tan ^ + 3 sin 0o(cot 02 — cot 0i) j e (3-30) Table 3-1 relates the various parameters of a fitted bearing with its performance, under the requirements of maximum load capacity for a given minimum film thickness. This shifts the position of the load vector arbitrarily along the bearing surface. The requirements for Table 3-1. Optimum Conditions in Fitted Bearings 0, deg 30 60 90 120 150 0i 129.1° 92.2° 64.3° 40.7° 19.4° a 17.5° 35.8° 55.7° 77.7° 101.1° a/P 0.58 0.596 0.62 0.65 0.674 hi/hmia 2.18 2.14 2.08 2.0 1.804 c/hm\a 2.18 2.14 2.31 3.02 5.43 h{)/hmia 1.4 1.48 1.62 1.84 2.131 0o 150.1° 136.4° 135.2° 142.4° 156.9° 4> i J 56.6° 38° 30° 28.4° 30.5° w uLNIt* 0.00451 0.01685 0.03309 0.04534 0.03924 _ ^min 0.0407 0.0774 0.1073 0.1284 0.1339 uLNR1 At 9.027 4.597 3.242 2.831 0.3413 ” min
Incompressible Lubrication; One-dimensional Bearings 53 minimum friction in the bearing are very close to those of Table 3-1. It should be noted that in fitted bearings the minimum film thickness is always at the outlet end. 3-5. Floating-ring Bearings. This kind of bearing has a thin ring float¬ ing freely between journal and bear¬ ing as shown in Fig. 3-10. The pur¬ pose of the ring is to reduce the shear losses by decreasing the relative speed between the mating surfaces. Since the losses vary as the square of the speed but vary only linearly with area, the rotation of the ring will always result in some reduction in drag, if all other parameters remain unaffected. It is clear that for equilibrium the sum of the two moments M i and M 2 acting on the floating ring must be zero. The expression for the drag is given by Eq. (3-12), which will be rewritten for our purposes in the following manner: Fig. 3-10. Floating-ring bearing. nULR 4x(l + 2e2) C (2 + «2)(1 — 4tvLRU [ 2 C [(l-€2)“ 2t nULR . (2 + €2)(1 - €2)*J (yjruLUe2 Cy/T=7* ^C(2 + e2)(l -e2)» By multiplying and dividing the last term by (R/C)( 1/e) and recalling the expression for W sin <t> given by Eq. (3-9), we have „ 2TpULR . cC w . * ~ C(1 - t2)» + 2R n * By a similar procedure the drag on a bearing is „ 2rgXJLR eC lir . 6 “ C( 1 - £2)* 2R Sln We thus have for the two moments (3-37) M, = - ^^L(RXN - R2Nr) Cl( 1 - €!2)* 4ir2nRzzLN R ^sin + Wi
54 Theory of Hydrodynamic Lubrication Since R\N — R2Nr « R2(N — Nr) and by writing W = 2LR2P and € = cos 6, we have by equating Mi to M2 „ _ . 1 - <.2)» v ^ - 1 + C2fl23(l - ‘22)k + P[C2t2(1 “ <s2)W + Cltl(1 - It is shown in Chap. 1 that the behavior of a unidirectionally loaded bearing in which both journal and bearing rotate in the same direction is the same as though the journal alone were to rotate at a speed equal to the sum of the journal and bearing speeds. Thus, according to Eq. (3-11) _ fRi\*n(NR + N) _ (2 + €l’).(l - €1*)K P 12t26i s2 = -(2+€22)(i -ef,)* 12t262 By using these relations to eliminate P in the expression for N, we have for the speed of the ring and for the relation between the two eccentricities - 3*i [e,(l - c,*)» + £-2 «2(1 - <2*)» j (3-39) The total power loss for a floating-ring bearing is " - [m; w”n*• *cfift)*<-u-v*\u <3-"» Figure 3-11 shows graphically the relation between the various ratios of clearance, radius, and speed. If one considers an equivalent standard bearing, denoted by subscript e, as having the dimensions Ne = N (\ = Cl Re = Ri = R then by using the more critical outer film as a basis for comparison, it results that the floating ring has a lower load capacity if Se/S2 > 1 and a higher capacity if Se/S2 < 1. The power losses are always less in the floating-ring bearing. 3-6. Porous Bearings. In a bearing made of a porous material, lubri¬ cant flows out of the bearing surface with a certain velocity VQ. If
Incompressible Lubrication; One-dimensional Bearings qy is the rate of flow per unit area, then oo dp q«=-Ty - = V0 v-o M where dp/dy is the pressure gradient at the bearing surface and $ is a property called permeability which varies with porosity and size of pores. Its dimensions are square inches. From the requirements of continuity we have for the porous matrix Vq = * V‘p = 0 so that., since ^ 0, ”o 0.2 0.4 06 0.8 1.0 1.2 1.4 1.6 1.8 2.0 y2p = q Rotio °f cleoronces, Ct/Ct The problem then is to solve Reyn- J'10-.3'1Var‘atiof.of rin« 8Peed with r , . bearing characteristics, olds equation for the pressures m the oil film simultaneously with that of Laplace for the porous matrix with a common dp/dy at the boundary, or *(»%)+* (V^\ = wdh 2*0j> I (3-41) dx\v dx/ dz\u dz) dx m dy |y=o d2p d2p d*p _ 3x2 ^ dy2 dz2 Two assumptions are made in solving this set of equations: 1. The bearing is infinitely short. 2. dp/dy is linear across the matrix and is zero at the outer surface of the porous bearing shell. Assumption 1 yields d2p/dx2 = 0; assumption 2 gives d2p dy2 Hence from Laplace’s equation = const = K d2p _ dz2 = -K and dp dy KH = -£?l dz2 H y = 0 OZ |y-0 where H is the wall thickness of the porous bearing. By using this expression in Eq. (3-41) with ^ ^ = 0, we have (A. + or d2p _ dz2 6m U e sin Q RC2 (1 + € cos ey + 12$>H/C r-TTii (8-42)
56 Theory of Hydrodynamic Lubrication Equation (3-42) solved and integrated for the resultant load along the line of centers yields W§ =M^V11±_* 2C2 ^ 6A.2 111 [fc2 - fc(l fc2 - k( 1 + €) + (1 + «2)] *) + (!- <)2 (1 + *: + «)* J (! + *-«) + (tan- - ‘I - tan- P-^ - -UP th? y/'& ( L fcV3 V3j L *:V3 vUJ/ (3-43) where k* = V2’t>H/C3. Figure 3-12 gives a comparison of the relative load capacities of porous Fig. 3-12. Load capacity of porous Fig. 3-13. Friction factor for porous bearings. bearings. and solid bearings both based on the short-bearing approximation. The coefficient of friction as given by -/ = n j 2ir2<S 1 + o sin CJ ' H -s/l - e ' 2 is plotted in Fig. 3-13 as a function of il, the relative load capacity of the porous and solid bearings. ONE-DIMENSIONAL THRUST BEARINGS The analysis of thrust bearings is made somewhat easier by the simplic¬ ity of the expressions for film thickness and by the less complicated bound¬ ary conditions. In most thrust bearings the film is nondiverging and continuous and the problem of negative pressures does not arise. The
Incompressible Lubrication; One-dimensional Bearings 57 pressures at both the inlet and outlet edges are simple boundary values, usually atmospheric. In fact, it can be shown that the elaborate con¬ dition p = dp/dx = 0 at the outlet cannot be satisfied for bearings having a converging film shape. A number of these one-dimensional thrust bearing solutions were obtained by Lord Rayleigh, who also derived parameters for optimum film shape and bearing proportion. The quantitative results here are no closer to reality than in the case of journal bearings, but, again, many of the dimensionless groups and much of the qualitative behavior deduced from one-dimensional analyses are also valid on the basis of more exact solutions. In reality a thrust bearing, which is designed to support axial instead of radial loads, looks like the device of Fig. 3-14. The popular simplifica¬ tion is to treat the sectorial segments of which such bearings consist as a simple plane slider as shown in Fig. 3-2. By using the system of coordi¬ nates given in Fig. 3-2, where U is negative with respect to x, Eq. (3-1) reads 6S| Fig. 3-14. Thrust bearing. <£ = -wh- ax — ho h3 From this with p(hi) = p(h2) = 0 Integration of the expression for load capacity by parts yields W-L^pto-L („ i;; - /; o, w-toVL(J*’-£ The center of pressure 1 can be obtained from The expression for shear stress is, from Eq. (l-16a), nU , h dp h 2 dx (3-44) (3-45) (3-46) ■dx (3-47)
58 Theory of Hydrodynamic Lubrication and by using the expression for dp/dx from above, we have , 4h — Sho h2 '■-'“(‘/Tt -“•£'!) (3-48) By using these general equations, a number of solutions for bearings having different film shapes can be obtained. 3-7. Plane Sliders. This bearing configuration is shown in Fig. 3-2, and its film thickness is given by , h2{a — 1) h = ax = ——~ x Jt> (3-49) where a = h\/h2. By integrating between X\ = hi/a and x2 = h2/af 0.20 0.18 Cf .0.16 Lp 0.14 0.12 0.10 /' 1.0 0.9 0.8. 0.7 0.6 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 a - hjht Fig. 3-15. Load capacity and friction in plane sliders. we have for the pressure distribution V = _ 6fiU (hi — ax)(ax — h2) (hi + h2)x2 h-l > x> -2 and for the parameters of maximum pressure t 2hih2 2d t 0 ” hi + h2 ~ T+~a k2 3nUa(a - l)2 2aa Integration of Eqs. (3-46) to (3-48) yields V o = ^ - tt1’ (rh)' ['” * - 2-r^r] - ~cPc' (^)B(a2 — 1 — 2a ln a) (a2 — 1) ln a — 2(a — l)2 = 2(a2 - 1) ln a - 3(a - l)2 ixUBL (3-50) (3-51) (3-52) (3-53) (3-54) (3-55) B 3(a + 1) ln a — 6(a — 1) h2 The dimensionless coefficients Cp and Cf as a function of a are plotted in Fig. 3-15.
Incompressible Lubrication; One-dimensional Bearings 59 By setting in Eq. (3-53) dW/da = 0, the value of a for maximum load capacity is obtained, and its value is5 a = 2.2. This is also evidenced from Fig. 3-15. Based on a = 2.2, the values of the other quantities are W = 0.1602 ho = 1.37hi F = 4.7 ^ h>2 Jo If the bearing is made up of a series of tapered-land segments, each segment being equal to B/n, while the total breadth is B, then which means that the bearing load capacity is (1/n) times smaller than that of an equivalent unbroken surface. This, of course, assumes that the taper over each segment B/n is equal to a. 3-8. Curved Sliders. If h = mxni then the value for the resultant load is _ 6pULB* I"(3n - 1 )(a-2+1'" - 1 )(a-3+2'" - 1) a~2+2'" - I] V(a1/n - l)2 L (2n - l)(a“*+l'")(3n - 2) 2(n - 1) J (3-57) By setting n = 2, that is, by using a parabolic surface and calculating its load capacity for the optimum value of a, which is 2.3, we have W = 0.163 If the film thickness is expressed exponentially, h = e?x, then by perform¬ ing the rather easy integration, we obtain t\r _ 3uUL Ta2 — 1 /3(a2 — a3)B~\ (0ah2)2 [ 6 ■*" a3 - 1 J or since = ln a, <-> For an a = 2.3, W assumes a maximum value of W = 0.165 h 2 The similarity between the load capacity for all these various bearings underlines the important fact that, once hi and h2 are fixed, the exact shape of the oil film is not of great importance. This conclusion is reached on the basis of two-dimensional analysis also.
60 Theory of Hydrodynamic Lubrication 3-9. Step Bearings. If in Eq. (3-1) we set h = const and consider isothermal conditions, the resultant pressures are zero; thus a bearing having a constant film thickness has no load capacity. However, if the film is parallel but has a step in it such as shown in Fig. 3-16, the bearing will develop hydrodynamic forces. The expressions for the film thick¬ ness here are By evaluating this expression for region Bi with the coordinate system Fig. 3-16. Step slider, as shown in Fig. 3-16 we have p = 0 for x — 0 V — Vo for x = Bi where pe is the common pressure at the step. The two conditions yield Similarly, for region B2 with the coordinates shifted to the leading edge, we have h = hi in region Bi h = h2 in region B2 By integrating Eq. (3-1) for a constant h, we obtain p = &pU^^x + Ci B Ci = 0 pc = 6pU^~^Bi C» = 0 Since pc is the same in both cases, we have or (3-59)
Incompressible Lubrication; One-dimensional Bearings 61 and the pressure profile is h2(Bih2* -j- B2hi*) P(I) " Ti?- / \ 6pU P(I) - T? 1 B,A,» + S2/n8 hi{Bih2* -f- B2hi*) B1A2* -I- B2hi* - 1 x for region £1 (3-60a) x for region £2 (3-606) which reduces to zero if hi — h2 or if B1 or £2 = 0. It is apparent from Eq. (3-1) by setting h equal to hi or h2 that the pressure gradient is dis¬ continuous at the step. The load capacity is „ = wt + w. = [*'xdx + W(h>-h2)L f*>xdx hi* Jo h2* Jo w 3pULBiB2B(hi — h2) ( . w B557(3'61) By writing £1 = £ — £2 and a = hi/h2j Eq. (3-61) can be rewritten _ 3pULBB2(B - B2)(a - 1) (Bja* + B — B2)h2i To find the optimum B2 and a we set dW n , dW n aF2 = 0 and 3a The first relation yields dW _ 3pULB(a — 1) dB2 W [*(£20* + B - £2)(£ - 2£2) - £2(£ - £2)(a3 - I)] _ n L (£2^ + £- £2)3 J " U and, to satisfy this, £2 - 2££2 + B22 _ /£ — £2V V ) £22 or £ = £2(a* + 1) The second relation yields dIF _ 3pULBB2(B - £2) I"£2a3 + £ - £2 - (a - l)3£2a2] da W [ (£*a* + £ - £2)2 J which yields £ = B 2 2a3 - 3a2 + 1
62 Theory of Hydrodynamic Lubrication By using the expression for B from above, B2 = £2(a* + 1) or 2a3 - 3a2 + 1 (a - l)(4a2 - 8a + 1) = 0 One root is a = 1 and the other one is a = 1 + V% = 1-866 The optimum value of B\/Bi is then §! = o2(2a - 3) = 2.549 r>2 With these optimum parameters, the performance of the step bearing is w 0.2052Mt/L£2 h2* B = 0.4262 / = 4.091 ^ It can be shown by the use of calcu¬ lus of variation that a stepped film is the optimum film shape for a slider bearing. This is true also for the more general case when the viscosity of the lubricant is considered to be a function of pressure. 3-10. Composite Bearings. The bearing termed “composite” is made up of a combination of tapered- and flat-land bearings, as shown in Fig. 3-17. Its film function is h = ax for region B\ Fig. 3-17. Composite slider. h = h2 for region B2 For region B2i taking the exit edge as the origin, we have for boundary conditions p = 0 at x = 0 p = pc at x = B2 with the common pressure pe given by Pc = 6/i t/f? 2 For Bi, writing a in terms of the taper hi — h2f > hi — h2
Incompressible Lubrication; One-dimensional Bearings 63 and by rewriting the expression p = f(x) into p = }{h) dp _ dp dh _ h\ — h2 dp dx dhdx Bi dh we have dp_^UBl(ha 1\ dh hi_htyh3 h,j and upon integration „ru\ SpUBifl h0\ n rW = h^h2{h-2h*) + C2 The boundary conditions for this equation are: p = p< = 6hUB2 at A = A2 hi p = 0 at h = hi By using these two conditions, we obtain p = 6»U |b2 (^-t - ^ ^ - 2^-^ j J (3-62) i. _ 2hih2(Bihi + B2hi) 0 (A, + A2)[B,A2 + 2B2A,V(Ai + A*)] ( ) It should be noted that the expression for pc is a function of the local h alone, and thus p will be continuous at the boundary. The total load is given by ■dh ri2 fB* fhl B i W = / Lpdx — / Lp yo fti — a2 _ 6Ml/LB!2 (B2 /B2 \ /A + 1 1 A + 1\ A2* (B,VBi /Vfc + 2 2 7* + 2/ -rh + i[^-m]| <"*> where k = ^-=A2 7 = r>7 ^/B')kik + D 2(Bi/Bi)(k+ 1)2+ (k + 2) With, --.</($-1) [(& - k) {; •"+j; (3 w - c) H l'“i2i:[3<,-’>rT5-2]-<h' _,[M- A: + 2 + 1) 6_ A* —|— 2 (3-65)
64 Theory of Hydrodynamic Lubrication If in the equation for ho we set B2 equal to zero, the expression reduces to that of an ordinary plane slider. With B2 > 0 the value of ho is less than for tapered-land bearings, which means that the point of maximum pressure has shifted closer to the exit edge. However, since ho > h2 this peak will always remain in the tapered portion of the bearing. The load capacity of a composite bearing is at an optimum when Bi/B2 = 5, at which point it will be 25 per cent higher than for an equivalent ta¬ pered-land bearing. Its losses, how¬ ever, will always be higher, as can be easily deduced from the longer span of its minimum film thickness h2. 3-11. Pivoted-shoe Bearings. The pivoted-shoe bearing is one which, instead of being rigidly fixed with respect to the runner, is sup¬ ported by a pivot and is thus free to assume any inclination. This it will do in a manner that will yield the highest load capacity. The per¬ formance of these bearings can be calculated from the results obtained for plane sliders. I n a pivoted-shoe bearing the resultant force obvi¬ ously has to act through the pivot, and this pivot position must coin¬ cide with the center of pressure £ derived for the tapered-land bear¬ ing. Each £ is associated with a certain minimum film thickness h2, and h2 in turn is related to load ca¬ pacity. Thus, from Eq. (3-54) [6) Pivot locotion Fig. 3-18. Performance of pivoted sliders. (H)#(a2 - 1 - 2a ln a) (a2 - 1) ln a - 2(a - l)2 For a given pivot position, a can be found from Eq. (3-54) and W from Eq. (3-53). Since these relationships are implicit in a, the results for W = /(a,£) are as given in graphical form in Fig. 3-18. The pressure distribution in plane sliders is always asymmetrical with respect to B/2. The resultant force will thus always be off center, and it follows that in order to have any load capacity, the pivot must be located at £ > B/2 from the inlet edge. It remains, however, a fact of
Incompressible Lubrication; One-dimensional Bearings 65 pivoted-bearing operation, both journal and thrust, that the bearing will operate satisfactorily with the pivot located in the center. Various explanations for this behavior may be offered, and some of the explana¬ tions are discussed later in the text. One of the more obvious reasons is that the surface of the pads is never flat both because of geometrical imperfections and because of the cantilever effect, which tends to bend the pad surface into a circular or parabolic shape. Fig. 3-19. Curved pivoted slider. In the following analysis it will be seen that a value of £/£ = % yields load capacities if the pads have some curvature. The equations are derived for a parabolic pad surface, but it can be shown that, since the change in elevation, hc of Fig. 3-19, is very small, a parabolic and a circular arc will yield similar values of hc as a function of x. By putting the coordinate axis at the leading edge, we can write he = hc — 4//c(* - y2y and for the film thickness h h = hi — x(hi — h2) — hc = h2{b[4:(x — %)2 — 1] + (a — ax + x)) where £ = x/B, b = He/h2, and a = hi/h2. By substituting h in Eq. (3-1), we get where the constants of integration can be evaluated from p(0) = p(l) = 0. The load capacity is given by
66 Theory of Hydrodynamic Lubrication W = B P pLdx = m^2L Ci. yo /i 2 (3-66) where Cp is a result of numerical integration and is given in Fig. 3-20. The load in Eq. (3-66) is given in terms of the outlet film h2 which, 0.66 Fig. 3-20. Load capacity of curved pivoted sliders. (a) Load coefficient based on minimum film thickness; (b) load coefficient based on outlet film thickness. Fig. 3-21. Pivot position for curved sliders. however, is usually not the point of minimum film thickness. The latter can be found from ^ = 8x6 - 46 - a + 1 =0 dx which yields for the minimum film thickness the following: - ®_-j m1 h ~h rfL+j _ - d2 _ j 8b + 2 2 [ 2 1C6 J »UB2L[a+\ (a- l)2 .I’ nUB*L _ [-2 m~~b\ Cp = -hi~Cpi Thus W = hi Cpi too is given in Fig. 3-20. The coefficients of the remaining parameters £ p nUBLn B Fi ~ -J— Cfl hm\n C f 1 f _ ,vmin i 3 ■ ~B C7i are all given in Figs. 3-21 and 3-22.
Incompressible Lubrication; One-dimensional Bearings 67 This analysis is an example of thrust bearings where negative pressures do occur. From hmin on, the clearance diverges and, depending on the value of He and other parameters, larger or smaller regions of negative pressure may appear. The integrations above did not exclude them, and thus the combination of pad curvature and negative pressures com¬ bined to give pivot positions on either side of the pad center. Over the range of f/B from 0.5 to 0.6 the optimum value for Hc/hmia is about 0.35. For this value of He/hmin the optimum pivot position is f/B = 0.55, and this yields a 10 per cent increase in load capacity as compared to an optimum flat-land pivoted bearing. Fig. 3-22. Friction in curved pivoted sliders. An additional advantage of this type of a bearing is the ease of starting; for even at rest a converging wedge exists between runner and leading half of the pad. Thus upon inception of motion, hydrodynamic pressures build up much faster than between flat surfaces. SOURCES 1. Sommerfeld, A.: Zur hydrodynamischen Theorie der Schmiermittelreibung, Z. Math. u. Physik, vol. 60, p. 97, 1904. 2. Sedney, R., A. Chames, and E. Saibel: The Reynolds Lubrication Equation with Smooth Outflow, Proc. First Natl. Congr. Appl. Mech. 3. Cameron, A., and Mrs. L. Wood: The Full Journal Bearing, Proc. Inst. Mech. Engrs. (London), vol. 161, p. 59, 1949. 4. Du Bo is, G. B., and F. W. Ocvirk: Analytical Derivation and Experimental Evaluation of Short Bearing Approximation for Full Journal Bearings, NACA Rept. 1157, 1953. 5. Lee, J. C.: Analysis of Partial Journal Bearings under Steady Load, ASME Paper 55-LUB-l, October, 1955. 6. Kingsbury, A.: Optimum Conditions in Journal Bearings, ASME Paper RP-54-7, 1931. 7. Morgan, T. V., and A. Cameron: Mechanism of Lubrication in Porous Metal Bearings, Conf. on Lubrication and Wear, Paper 89, London, 1957. 8. Rayleigh, Lord: Notes on the Theory of Lubrication, Phil. Mag. vol. 35, no. 1, 1918. 9. Abramovitz, S.: Theory for a Slider Bearing with a Convex Pad Surface; Side Flow Neglected, J. Franklin Inst., vol. 259, no. 3, 1955. 10. Shaw, M. C., and E. F. Macks: “Analysis and Lubrication of Bearings,” McGraw-Hill Book Company, Inc., New York, 1949.
Basic Differential Equations 15 Control volumes such as illustrated in Fig. 1-5 are imaginary volumes generally fixed in space through which the fluid at continuously varying velocity, temperature, pressure, density, and viscosity is allowed to pass. Since an energy equation is desired, all the component energies will be summed over this volume for a unit interval of time according to the first law of thermodynamics: Ei + Hdo — E, + E0 + Hdb where Ei = energy transported into the control volume E0 = energy transported out of the control volume E, — energy stored transiently in the control volume Hdo ~ work done on the fluid volume by the surroundings Hdb = work done by the fluid volume on the surroundings Steady-state conditions are assumed, so that the above equation becomes E0 — Ei = Hdo — Hdb (1-19) There are two modes in which energy may be transported into and out of control volumes: by conduction according to Fourier’s law and by convection of intrinsic energy, i.e., transport of fluid possessing kinetic energy and internal energy. A possible third mode, radiation, is neglected. The other energies involved in the energy balance are the mechanical works done by the surface stresses and body forces through an incremental distance in an increment of time. For the lubrication problem at hand, body forces, such as gravity, are neglected. The transported energies and mechanical works involved are indicated separately in the control volumes of Fig. l-6a and 6. So as not to encum¬ ber the sketches, not all component energies are indicated. It is to be noted that differential changes in energies are taken about the mid-point 0 in the control volumes. The transported energies of Fig. l-6a summed over the surfaces of the control volume according to the left-hand side of Eq. (1-19) are *• - * - (pfc*+:V+“E*] -' [£(*£)■+ «(*S) where the intrinsic energy is given by e = *±4±^ + Jc.T The mechanical works indicated in Fig. 1-66 must also be summed over the volume surfaces. However, an interpretation of what is meant
16 Theory of Hydrodynamic Lubrication by work done on and work done by a volume in terms of the surface stresses and fluid velocity is first needed. All the works done by the fluid volume are on the upstream surfaces of the control volume, i.e., where the velocity components are in the opposite direction to the stress dog A/ E--q dA etc do/ A/ dir A/ °> + 17 ~'*+77T dT/X A/ dr/ A/ T"+iJ/ T’*'+IrT . du A/ „JjLki7 > it 2 \/ dr A/ d<rr A/ dK A/ U) Fig. 1-6. Control volumes, (a) Transported energies; (b) mechanical energies. components. With this viewpoint in mind, the right-hand side of Eq. 1-19 becomes d HJo Hdb — (llOz “f" tJTU* "f" dy d\ , (UT V<7y -j- WTZy) a,) j Ax Ay Az By equating the expressions for E and H according to Eq. (1-19), + ^ {utxz -j- vrUi + W0z] |~d(pue) d(pve) d(pwe) [ dx dy dz
Basic Differential Equations and by rearranging some of the terms. 17 [*S-S+-£]-'[£(*SK(*S)+£(*S)] -*(£+»+£)+*(£+&+&) (&+1+b)+('• B+- S+S)+- (w+ S) (I+Ij)+'"(b + £) <■-“> + “Tyi where the first parenthesis term of Eq. (l-20a) reduces to its stated form because of the continuity equation d(pu) d(pv) d(pw) _ dx ^ dy ^ dz The equilibrium equations of fluid flow for steady-state conditions (dv/dt = 0) and zero body force are given by the expressions preceding Eqs. (1-1), namely / dll du : du\ dox drzy p{UYx + Vd-y + WYz) = -dI + ^ ( dv . dv . di>\ drux p\UTx + Vd-y + Wdi) = l>x ( dw dw dw\ drtz drty , do- p\udl + vdj +wdl) = aT + ^ + aF dTx- dz i doy . dryx ^ dy ^ dz The substitution of the expressions for o and r into the last three equa¬ tions yielded for us previously the Navier-Stokes equations. Substitut¬ ing the last three equations and equations for o and r into Eq. (1-206) yields ( de . de . de\ . I" d /, dT\ . d (, dT\ . d /. dT\ 1 p\uTx + vd-y + wTz) ~ ,/[aiV+ 9v\ dy) + d’z\!fe = P (u d[{u2 + v2 + w2)/2] v d[(u2 + »2_±w2)/2] dx d[(u* + ; dy ^ dz j p\dx^ dy ^ dz From e = (u2 + v2 + w2)/2 + JcvT we have jp r u +v ^p.+w a(c’7’) dx dy dz 1 , [du di> dw + ~dz + (1-21)
18 Theory of Hydrodynamic Lubrication where - * [* (£)‘+’ (£)'+*(£)■- i (£+S+£)' . (du dv\2 . (dv . dw\2 . fdw dw\2l + + ^ + + ^ + a?jJ It is to be noted that only a d{cvT)/dt need be added within the first bracket of Eq. (1-21) to make the energy equation applicable also to tran¬ sient states, subject only to the limitation that the flow be laminar. The first bracket term in Eq. (1-21) is the convection of internal energy of the fluid. The second bracket term is the rate of work done by a differential volume of fluid in expansion against the surrounding pressure. fro dhdh -f. Fig. 1-7. Energies at solid boundaries with slope, (a) Transported energies; (6) conducted heat. The third bracket term is the rate of heat conduction in the fluid. And the fourth term is the rate at which kinetic energy is dissipated into heat. A convenient starting point in integration of the energy equation across the film thickness is at Eq. (l-20a). Thus, Eq. (l-20a) becomes /:[ d(pue) d(pwe) 1 , |A =/:[ d d (iKTz + VTvx + WTZX) -f- ~ (UT -1- VTyx + w<Tt) dy “1“ (WTXy “I- V(Ty -f" WTZy) For the moment the terms of interest in the above equation are: pve kd-T dy (UTzy + V<Ty + WTzy) which must be investigated at an incremental boundary element such as illustrated in Fig. l-7a. An energy balance is taken on Fig. l-7a for
the first term which becomes + pwe + Basic Differential Equations (Axo “1“ fozo AZ| Kyi Ezi)h = 0 dh (Az)2 ~dz 2 d(pwe) A_~| \dh Ax dh A ] A du (Ax)2 19 d(pu«) “If aA d6 Az] pue + ~&rAx\ [aiAx + TzT\ * ~pue . 1 [ dh Ax . dh . 1 . du (Ax)2 Az\ [d5T + ^A2JAx-pwa5 — dz | | dx 2 1 dz | 1 dx 2 — pve Ax Az = 0 where average incremental heights have been used, and hence dh dh pve = pue — h dx h = pWe~dz In lubrication problems, the boundaries are usually such that and thus u = w =0 I h |A pve\ =0 If the fluid has intrinsic energy at this boundary, then v ^ = 0; this may be verified by a mass balance of the same form as used above. At the zero-slope boundary pve = 0 The heat conduction term may be evaluated from Fig. 1-76 as - K,(T - 7V) (1-22) dy _ dTdh dx dx , , dTdh + k-r- — h dz dz where the approximation, Jr.o V1 + (^) V1 + (S) - has been made. Tw is the stator-plate temperature, which may or may not be a function of x and z, and K\ is the heat transfer coefficient at the fluid-solid interface. If the runner has the same temperature distribu¬ tion as the stator plate, then dT k dy K2(T - Tw) The surface mechanical works are (urxy + V(Ty -I- WTzy) = 0 (UTXy -{” V(Jy -j“ WTZy) 0 TXJ
20 Theory of Hydrodynamic Lubrication since u = v = w = Ow = U, and v = w =0. \h |/» \h |o ’ |o |o Substitution of the above equations into the integrated form of Eq. (l-20a) yields jl [ + hI""> ] d> - J /„' [k (‘ H) +1 (* S) ] *• -j(t££+ti£i£)l+K’<T-T-> ~ Jo ^U(Tz VTyX ^UTlZ ^ ~~ ^Txv o (l-23a) where Kt = Ki + K2. Carrying out the same operation as in Eqs. (l-20a) and (1-206) gives for the right-hand side of (l-23a): Right-hand side iide = p ju d[(u2 + v2 + w2)/2] dx + v d[(u2 + v2 + w2)/2] + ^ d[(u2 + v2 + w2)/2] | dy dz | fh (du dw\ , fh ( d [ (du dv\ -Jo p(di+^)dy-Jo |MapK^ + ^/. . d ( di/\ dp 2d (du dv dw\ L ^ 3 ^ M (to + ^ + Tz) + W Fy [" (If + %)]) dy + lo *" ^ - C/T- lo (U236) where *" - - [*(£)'■+2 (£)'-l(s + r» + S) (I + S) /dw diA2 I I ^ _L (Urn ■ ^1 ^ds: dx ,/ * ^ + dx/ dx + \dl/ + dz) dz\ + 1 And now making the approximation that since the film thickness h <$C B, L, then v « 0 and T, p j* f(y). In addition, then p and p are inde¬ pendent of y. Hence the first integral in Eq. (l-23a) may be reduced, since the continuity equation becomes d(pu) d(pw) = 0 dx dz and * = + Jc.T Also, the second integral in Eq. (l-23a) may be integrated directly. The first integral in Eq. (1-236) will cancel (for v = 0) a like integral on the left-hand side of Eq. (l-23a). With these approximations, equa¬ tion (l-23a) reduces to
Basic Differential Equations 21 + K,(T-T.)\ --, £(£ + %) i, -' f; (■ w+• ®o* - ‘u 11.+r ■" v-2w where ,, f0 /du\2 (dw\2 2 (du dw\2 (du dw\2l . OA,* * =m[2U/ +2w _n^ + */ +U+^jj (U24b) which cannot be deduced from Eq. (1-21) by stating v « 0. The next step in the reduction of Eq. (l-24a) is the neglect of any terms of insignificant size. Dealing first with terms of a mechanical nature, one of the expansion terms is by requirements of continuity, du n / p dp\ pei = 0{u7^) Hence the ratio of one of the largest viscosity terms to the expansion term is pu d2u/dy2 p du/dx W pw) which is approximately equal to 1 X 103. By taking a similar ratio with one of the terms in Eq. (1-246), uu d2u/dy2 _ n /£2\ 2u(du/dx)2 \h2) which is nearly equal to 2 X 107. It might be noted that °(M/o*uSdi/)=0(^SL) Substitution of the relations for the velocity as given by Eqs. (1-5) U = - yh) + U w = ^(y* - yh) (‘-D into Eq. (l-24a) and using the above approximations yields ([(pVh h> dp\ d(c,T) h3 dpd(c,T)~\ tL\ 2 12k dx) dx 12v dz dz J -[l(tk^) + i(hk^}} + K^T-T->)
22 Theory of Hydrodynamic Lubrication For ordinary lubricants, the characteristic values of the parameters are such that conduction is a minor mode of heat transfer. Further, the heat transfer coefficient at the fluid boundaries as reported in Ref. 5 is very small. Since it is expected that the fluid temperature rise through a bearing will not be large, the specific heat of the fluid will essentially be constant. With the above-mentioned fluid parameter characteristics, Eq. (1-25) becomes r ttt . l. T/i W dp\dT h2 dp dT~] _ \2nU> \, , /t4 (inUdx)dx (i/it/ dz dz J h ( + 12M2t/2 [(g)’ + ©II <■*> Equation (1-26) states that all of the heat generated within the fluid because of viscosity is carried away by the mass transfer of the fluid and that no heat is gained or lost through the bearing surfaces. Equation (1-26) can be made more convenient for numerical work by reducing it to nondimensional form. By setting x = x/B, z = z/Bt h = h/B, /z = m/mi, p = p/pi, V = pB/GmU, T = TpiJcvB/pJJ, where B is a representative length in the x direction, Eq. (1-26) becomes -h(i _ _ M3 dp ar = 9 m 11 , 3h4 r/ap\2 /dpV]) a dxjdx a dz dz a I P L\d*/ \di) J) (l-27a) Likewise, the Reynolds equation becomes 1-5. Equation of State. In Eq. (1-26), m is a known function of p and T and h a known function of x and z. One more equation is needed because there are three unknowns (p, p, and T), and it is provided by the equation of state given by pv = (5iT (1-28) For lubrication with a gas, the assumption that the gas obeys the perfect- gas law will be adequate. Since p and f are quantities of practical importance, the procedure might be to eliminate p from (l-27a) and (1-275) by the substitution p — 1/6(n — 1 )pT. For lubrication with a liquid film the choice of an equation of state is more difficult. Even for simple liquids the equations proposed are modified van der Waals type of considerable algebraic complexity, so that their introduction into Eqs. (1-27) would increase the complications to such an extent that the solution would probably be a major computing operation. If it be sufficiently accurate to ignore the variation of p, /z, etc., with temperature, or if the variation with temperature can be replaced by a
Basic Differential Equations 23 variation, known a priori, with x and z, then Eq. (1-276) becomes an equation for p only. The solution thus obtained can be inserted into Eq. (l-27a), which then becomes an equation in T only. This situation, which also arises in other branches of applied mechanics, enables the equations to be solved successively instead of simultaneously, but only in the order (1-276) to (l-27a). However, as soon as it becomes necessary to take variation with temperature into account, the equations become interlocked and must be solved simultaneously. SOURCES 1. Reynolds, O.: On the Theory of Lubrication and its Application to Mr. Beau¬ champ Tower’s Experiments, Phil. Trans. Roy. Soc., London, vol. 177, part 1, 1886. 2. Weilder, S. E.: Data Folder DF-54-AD-7, General Electric Company. 3. Goldstein, S.: “Modern Developments in Fluid Dynamics,” vols. I and II, Oxford University Press, New York, 1950. 4. Vogelpohl, G.: Heat Transfer in a Bearing from the Lubricant to the Gliding Surfaces, VDI-Forschungsheft, July-August, ed. B, vol. 16, no. 425, pp. 1-26, 1949. 5. Cope, W. F.: The .Hydrodynamic Theory of Film Lubrication, Proc. Roy. Soc. (London), A, vol. 197, p. 201, 1949. 6. Sternlicht, B.: Energy and Reynolds Considerations in Thrust-bearing Analysis, Proceedings of the Conference on Lubrication and Wear, I.M.E., 1957, pp. 28-38. 7. Pinkus, O.: “Counterrotating Journal Bearings,” General Electric Internal Publication R60MSD322.
CHAPTER 2 HYDRODYNAMICS OF SIMPLE CONFIGURATIONS It is not the purpose of this chapter to deal with any general problems, nor even with any specific branch of fluid flow. The subjects were chosen solely on the basis of whatever relation they may have to the study of bearings, be it as actual components of lubrication systems or as an idealization of a complex bearing geometry. Thus capillary tubes and orifices are an integral part of hydrostatic bearings, and their character¬ istics must be known before the performance of such bearings can be calculated; and the study of flow in concentric and eccentric cylinders is applicable to journal bearings and may throw some light on the flow of lubricant in the clearance space. GENERAL EQUATIONS OF MOTION FOR COMPRESSIBLE FLUIDS The three forces affecting the flow of a gas in a slot are the pressure force, the viscous force, and the force required to accelerate or decelerate the fluid. If the depth of the slot is very small and the viscous and pressure forces are very large in comparison to the inertia force, then the viscous force can be equated to the pressure force. For laminar flow this results in a simple theoretical equation relating pressure distribution, temperature, mass flow, and slot geometry. For turbulent flow use is made of the concept of resistance X, where * = \7~~T~ (2-1) Mpul, t being the “skin friction” per unit of surface area in contact with the fluid and wmvg the spatial mass velocity. The relationship between X and the Reynolds number Re, based on the hydraulic mean depth for turbulent flow, is given empirically by Blausius1-* as X - °*079 (2 2^ Xr - Ri« (2'2) and this value can be used to develop an equation corresponding to that * Such superscript figures indicate references listed under Sources at the end of the chapter. ‘24
Hydrodynamics of Simple Configurations 25 obtained theoretically for laminar flow. The assumptions made in the derivation of the flow equations are the same in both cases. The force required to accelerate or decelerate the fluid is assumed to be negligible; the pressure distribution over any cross section is assumed to be constant; and the temperature of the gas as it flows along the slot is assumed to be constant or a known function. 2-1. The Theoretical Equation for Laminar Flow. With the condi¬ tions as stated above, we can start with Eq. (l-5a). For the flow of a fluid between stationary walls we have Ui = U2 = 0, and Eq. (l-5a) becomes — hy dp u = 2/x dx giving a parabolic velocity profile. The mean velocity at any cross section is Wav 1 [h A = hJoUdy=- dx 12/x (2-3) (2-4) If the width of the slot is denoted by 6, where 6 is a function of x, the weight flow along the slot is then given by G = pgbhua, — _ &P ^ 12/i dx from which we obtain dp = _ 12 pG dx gpbh3 and by substituting in this equation the value gp = p/<HT, we obtain pdp = 12/i(R TG bh3 dx (2-5) Fig. 2-1. Elementary volume between two plates. If T, b} and h are known functions of x, this equation can be integrated and the theoretical pressure distri¬ bution so obtained. 2-2. The Empirical Equation for Turbulent Flow. Figure 2-1 shows an elementary volume of fluid between two plates, its length in the x direction being infinitesimally small and its width in the z direction being unity. The force F resisting the motion of the fluid may be considered as due to the friction acting at the surface of contact between the fluid
26 Theory of Hydrodynamic Lubrication and the wall. If the skin friction per unit area of wetted surface be denoted by r, then Fi = rAB cos 6 + rDC cos <f> that is Fi = 2r dx The pressure force acting in the direction of motion is given by ^2 = M ^ dx^j (AB sin 0 + CD sin <f>) ~(P + tx dx)(h+Txdx) = pk + (p + \txdx) Txdx - (P + txdx) (h + Txdx) which, neglecting second-order terms, reduces to F2 = -h^dx dx By equating the two forces Fi and F2, we have dp _ _ 2r dx h By substituting Eq. (2-1) we get ~ = - —(2-6) dx h The coefficient X is dependent on the Reynolds number Re. For laminar flow the value of Xl is x =24 L Re On substituting \L in Eq. (2-6), we have dp _ _ 24 uJVJ = 24 __GP_ dx Rep h Re g2pb2hz where Re = p gbp dp _ 12 pG Hence dx bhzpg , \2p(RTG , fo or pdp (2‘7) which is the same as Eq. (2-5). To obtain the corresponding equation for turbulent flow, we substitute in Eq. (2-6) the value \t
Hydrodynamics of Simple Configurations 27 0.079 G2 dx Re* h Re* g2Pb2h3 0.067m*G* hW'pg* , 0.067m*(RTO* , 0, pdp = Piy—(2'8) FLOW THROUGH NARROW SLOTS The general equations (2-7) and (2-8) can be integrated if b, h, and T are known as functions of x. We shall now apply these equations to various specific cases. 2-3. Isothermal Flow. Constant Area Slot. For a constant area we have h = const and b = const, and integration of (2-7) gives for steady laminar flow 2 2 24n&TG . . fo m Pi2 - gp— (*2 - *i) (2-9) On integrating Eq. (2-8), we obtain for turbulent flow 0.133„*(RTO»(*. - X,) P> “ P2 = pwp ( ) Diverging Width. Let us denote the rate of width divergence by b — ax. Then if the frictional effect of the two side walls is neglected, the flow is analogous to the radial flow between two circular flat plates of constant film thickness, and by symmetry the pressure at any given radius is constant. Provided, therefore, that the divergence is small, the pressure in any plane perpendicular to the axis is approximately constant, and Eqs. (2-7) and (2-8) may be applied. By substituting in these equations b = ax, we obtain for laminar flow , 12 »<RTG dx pdp — t- ahz x which when integrated becomes 2 2 24M(Rrci x2 /011N (2'n) and for turbulent flow , 0.067m*<R7’G* dx pdp ^ which when integrated becomes \Xi« Xi’V . 0.178x xi P* P2 ct/'h3gii I •*••* ) (2-12)
28 Theory of Hydrodynamic Lubrication Diverging Depth. Let us here denote by h = fix (b = const) the divergence of slot depth. Because the divergence of the two plane sur¬ faces is very small, the component of velocity perpendicular to the axis of the slot must also be very small, and the pressure distribution over any cross section is, therefore, assumed to be constant. For laminar flow, substitution of h = fix in Eq. (2-7) yields , 12 dx Vdp and, when integrated, this becomes Pi2 - p2 W3 \Xi2 X22 By the same substitution in Eq. (2-8), we have for turbulent flow 0M7n*(RTGK dx (2-13) p dp = — which when integrated gives blip SgK t 0.067MM<nrow / 1 1\ /01,, Pl -w ~ &) ( } 2-4. Flow through Orifices in Series. Let us consider two identical i ft, Pbi | | T Po^ai | 1 \h 1 ..I Lj Ps 1 Pb o b (<7) \b\ Fig. 2-2. Orifices in series. orifices a and b as shown in Fig. 2-2o. The pressure ratio across a and b is „ _ Pal ^ _ Pb2 ra Tb Pal Pb1 where pb i = pa 2 The flow for a single orifice based on the perfect-gas equation is given by G = CA [2g (^j (r2/* - j* (2-15) where vt — specific volume at supply condition k = isentropic expansion coefficient From continuity of weight flow we have Ga = Gb. Thus, by assuming that the areas and discharge coefficients are equal and expressing the
Hydrodynamics of Simple Configurations weight flow by Eq. (2-15), we have TaW{ 1 - ris/*(i _ rt<i-*)/*) Val Va2 (2-16) 29 If isothermal flow is assumed, we get Vg2 _ Pal Va\ Pa2 On rearranging the terms and substituting the above equation into Eq. (2-16), we have By knowing the pressure ratio across one orifice, the pressure ratio across the other orifice can be found from Eq. (2-17). It can be easily shown (Fig. 2-26) that, for small area ratios, the velocity at the last orifice is given by If the area and discharge coefficient of each orifice are given by A and C, respectively, then the volumetric flow through the last orifice downstream is given by If the pressure ratio across the last orifice is equal to or less than the critical • (0.53 approximately), then the volumetric flow is a constant and is given by The flow through the next orifice upstream (which is not choked) is given by Eq. (2-15) and is or Qi = CAUi Qi = CA lig ptf! T1 - Jh r = 2? < 0.53 Gi = gpiQi = gpiCA 2g - (0.53)<*-»'*]|H (2-18)
30 Theory of Hydrodynamic Lubrication If we equate the weight flows through the orifices and still assume that the areas and discharge coefficients are equal, then [ 1 - (0.53)(t-l)/‘l = ^ (r22'* - r2<*+>«‘) Vl L J v2 Assuming again that pv = constant, the above equation reduces to r2_2(r22/* - r2«+l)lk) = 1 - (O.W'~»lk (2-20) From Eq. (2-20) we see that the pressure ratio r2, and therefore any pressure ratios upstream, will be constant. Therefore, the flow from the inlet up to the orifice with the pressure ratio r2 will be choked. Under Table 2-1 Number of orifices in series Approximate critical pressure ratio Critical supply pressure (exhaust to atmosphere), psia 1 0.528 27.8 2 0.430 34.2 3 0.370 39.8 4 0.330 44.5 5 0.305 48.2 6 0.290 50.7 7 0.275 53.5 8 0.265 55.5 9 0.255 57.7 10 0.250 58.8 these conditions the pressure upstream of the last orifice is directly pro¬ portional to the supply pressure. Looking at Eq. (2-18) we see that the weight flow through the last orifice is equal to gpi(piVi)** times a constant factor or Gi — gpi(piVi)^K K - const Now, since gpi = \/v\, the above equation reduces to «■ - {v)“K From the perfect-gas law we have = pi Thus we get a - ”• (irrl)"x
Hydrodynamics of Simple Configurations 31 Since pv = const, GiT = const and we have G\ = piKi K i = const We have already deduced that pi/p, = const; therefore, the equation above becomes G\ = p,K2 — < critical volume K2 = const P* and we see that the weight flow through a series of orifices with the exhaust orifice flow choked is directly proportional to the supply pressure, provided that the discharge pressure is constant. For a discharge to the atmosphere, the critical pressure ratio and critical pressures are given in Table 2-1. INCOMPRESSIBLE FLOW 2-6. Flow between Parallel Walls. The simple case of laminar flow of an incompressible fluid between two parallel surfaces of infinite extent is given by Eq. (l-5a). By placing the coordinate axis y — 0 at the center of the film, i.e., at h/2, and setting U2 = 0, we obtain -?(• + *!)-si [>-*©'] For the case U = 0, that is, when both surfaces are at rest, Eq. (2-21) gives a parabolic velocity profile; for dp/dx = 0 and U ^ 0 the velocity profile is linear. For the general case the velocity profile is made up of both contributions, and, depending on the value of dp/dx, the profile may be convex or concave, or it may even reverse itself. The point at which flow will occur in a direction opposite to U can be calculated by letting du/dy = 0 at the stationary wall. The result is that, for dp/dx > 2 Up/h2, the flow will reverse itself over certain values of y. The general behavior of the velocity profile is shown schematically in Fig. 2-3. If the two surfaces —h/2 and +h/2 are kept at uniform temperatures of T2 and T\, respectively, then, since dT/dx = 0, we have from the energy equation . d2T (du\ dy1 M \dy) By using du/dy from Eq. (2-21), d*T _ IPp [ 16umtkXy (twyVl 'dy2 h2k L Uh ^ \ Uh J J where umAX = — ^ is the maximum velocity. 8p dx
32 Theory of Hydrodynamic Lubrication The solution of the above equation is = t, + (p- - T 2 . fiUUn 3 k 1 + 2 fiUum 3 k The temperature gradient is dT = Ti - T2 , 2mUum dy h 3kh [i+80D1+tHi _ i6©' IS 2hk I \h) (2-22) + 16 um ~U \h) , 128wL 3 U2 \h (2-23) and depending upon whether the first right-hand term of (2-23) is greater or smaller than the remaining two terms, heat will either flow into or out of the upper wall. T->0 dx du I dy I > o ± = o dp dx dx dx du = 0 1 h £1 >° dy | h du dy dy | 1 * 2 2 2 < 0 > 0 2-6. Circumferential Flow between Concentric Cylinders. The lam¬ inar flow between concentric cylinders extending to infinity in the axial direction yields from physical considerations u '= u(r) v = w — 0 p = p(r) When these considerations are used in the Navier-Stokes equations, the following two differential equations are obtained dr2 ^ dr \r) pu2 _ dp r dr (2-24 a) (2-24 b)
Hydrodynamics of Simple Configurations 33 By using the boundary conditions of Fig. 2-4, we obtain by integration of Eq. (2-24a) U = [r(u>2ft22 - Ulft,2) - («2 - «,)] (2-25) and for the pressure distribution from Eqs. (2-25) and (2-246) * = PI + (g,»-B1y[W - — 2i?l2/?22(&>2 — <*>i) (o^fi^2 — Wj/?12) In til 2. where pi is the pressure at Ri. -W«i« fi2‘(«. - »,)* ^ f i)] (2-: 26) Fig. 2-4. Notation for concentric cylinders. If the inner cylinder is kept at rest, the moment of the fluid on a length L of the outer cylinder is M 2 = 2ttR22Lt2 — 2tR'?Lp,t «GD or Af, = 4irL “2 (2-27) This last equation can be used to determine m by merely measuring the moment M2. When only one cylinder rotates in an infinite fluid, we have for o>2 = 0, T — oo /?l200l w = r Af = ^tcpLR\2<j)i The energy equation for our case is H / dT\ = _ m /dw _ w\* r dr \ dr / A; \dr r /
34 Theory of Hydrodynamic Lubrication which gives the following temperature distribution T—T I M ^i4^24(o>i — 2)2/ 1 1\ 1_1"/b (RJ-RfY \RS r2/ T T fl Rl4R24(d)l-U)2)2 ( 1 l\ l2^k (RS-RW \RS RS), r ln Rt/Rx Ri (2“28) 2-7. Axial Flow in Cylinders. Concentric Cylinders. The differential equation for the axial flow of fluid in an annular space is + <2'29> By using the boundary conditions of w = 0 at Ri and R2i where Ri and R2 are the inner and outer radii, the velocity profile in an annular slot becomes + (2-M> By integrating w between Ri and R2, the flow becomes «-*■•>[*1’-ft’-ETIGTO] <M1> By setting Ri = 0 in Eq. (2-30), which gives the flow in annular slot, we obtain the flow of a liquid in a circular cylinder. The velocity profile is given by (2-32> 1 2 w-“"i»TzR> The velocity profile is parabolical, therefore 1 1 dp p 2 W.v, - 2 W,„ gM dz Ri The volume flow is given by Q = RS - ^ (2-33) The energy equation is given by . /d2T , 1 dT\ /dw\2 \dr2 r dr) M \dr) so that for a temperature of T2 at R2 and at r = 0 we have T-Tt+ (Ti _ T,) [ 1 - (2-34)
Hydrodynamics of Simple Configurations 35 Equation (2-33) is useful in the calculation of the performance of inter¬ nally pressurized bearings where lubricant is admitted through a series of capillary tubes. Flow through Eccentric Cylinders. If the cylinders are not concen¬ tric, the slot height h is a function of 6 and is given by the equation C(1 + c cos 6) (Chap. 3). If we are allowed to simulate the case of eccentric cylinders by two nonparallel developed surfaces, then the velocity from Eq. (1-56) is given by and now is a function of both coordinates. The flow is, by integrating w, For c = 0, the above equation reduces to (1-56). The exact analytical treatment of flow in eccentric cylinders is compli¬ cated. It represents a two-dimensional problem which is described by Poisson’s equation where w — 0 at the boundaries. A partial solution of this problem is given in Ref. 5, where the velocity profile is expressed in series form by and the various terms are, referring to Fig. 2-5, defined as follows: C = distance from the origin (z = 0) to either pole of a bipolar coordi¬ nate system w = - ^ y[y - C(1 + € cos 0)] (2-35) Q = (1 + ««*) (2-36) d2w dhv _ I dp dx2 dy2 p dz (2-37) where C + z p = q~zt~z 2 a variable in the complex plane
36 Theory of Hydrodynamic Lubrication An interesting result reached by evaluation of Eq. (2-37) is that the velocity distribution at any section in the variable-height annulus is very close to the one that would have resulted from a concentric case with the clearance equal to the dimension of the particular section under consideration. Equation (2-35) is thus a good approximation of the actual velocity profile. SOURCES 1. Piercy, N. A. V.: Aerodynamics, Elektrotech. u. Physik., p. 278, 1937. 2. Shires, G. L.: The Viscid Flow of Air in a Narrow Slot, ARC Tech. Rept., Cp 13 (12329). 3. Robinson, G. S. L.: Flow of a Compressible Fluid through a Series of Identical Orifices, ASME Paper 48-APM-4. 4. Shih-I-Pai: “Viscous Flow Theory,” vol. I, D. Van Nostrand Company, Inc., Princeton, N.J., 1956. 5. Poritsky, H., and Fend, F. A.: Laminar Incompressible Flow between Non-con- centric Circular Cylinders, TIS Repl., 57GL54, General Electric Company.
CHAPTER 3 INCOMPRESSIBLE LUBRICATION; ONE-DIMENSIONAL BEARINGS With the exception of Chap. 6 we shall be dealing throughout this book with hydrodynamic lubrication. By “hydrodynamic lubrication” we mean a process in which two surfaces, moving at some relative velocity with respect to each other, are separated by a fluid film in which forces are generated by virtue of that relative motion only. As in all other problems in engineering, the solutions on the following pages are based on certain assumptions, and in order to appreciate the degree of applica¬ bility of these results, a realistic picture of bearing operation will first be given. THE REAL BEARING Figure 3-1 shows a journal bearing operating with an external load W and speed (7. Under the physical conditions imposed, the journal will run at some eccentricity e, the region below the line of centers 00' form¬ ing a converging and the region above 00' a diverging space. From A to D the lubricant is being pumped by the journal into an ever-decreasing space with the result of building up high pressures in the fluid. The eccentricity e and the magnitude and distribution of these pressures will be such as to yield a resultant force equal and vectorially opposite to W. In the process the fluid is being continuously squeezed out the ends of the bearing, and this side leakage, plus any conduction and radiation that may exist, carries away the heat generated by the rotating journal. New lubricant is being delivered at point B to replenish the amount lost by side leakage. The hydrodynamics of thrust bearings are essentially represented by Fig. 3-2. Available bearing solutions even in their elementary form satisfy the basic requirements of continuity and momentum and express the per¬ formance of bearings as a function of load, speed, viscosity, and bearing dimensions. However, there are other significant features which are often disregarded either because of incomplete knowledge or because of mathe¬ matical difficulties. Together with the basic theory they constitute the physical reality of journal and thrust bearings. These features are: 37
38 Theory of Hydrodynamic Lubrication 1. Boundary Conditions. The pressure profile in journal bearings starts at the point where lubricant is admitted if 0i > 0 and at 0 = 0 is 0i < 0 ( —ir < 0 < ir). In a full bearing the pressure profile ends beyond hmiD, at 02, where it falls to a value very slightly below that pre¬ vailing at the bearing sides and then rises again to equal the boundary pressures. In those partial bearings where the arc ends before E and in most thrust bearings the pressure profile ends at the exit, much as shown in Fig. 3-2. The dip at the end of the pressure wave, region EG} Fig. 3-1. Dynamics of a full journal bearing. cannot be eliminated by simply raising the inlet pressure at B. Com¬ pared with the pressures prevailing from B to E, the values in region EG are negligible. 2. Striation. The fluid film in full journal bearings is rarely complete. If a deep axial groove is cut at B, the full film will start along B. If there is only a hole for admitting the lubricant, a full film will form along the dashed line B' of Fig. 3-Id. Between B and C the flow, because of the unfavorable pressure gradients, will consist of the shear flow less the pres¬ sure flow. At C, dp/dd = 0, and only shear flow prevails. From C to E, the flow consists of shear flow plus the pressure flow. Past hmin the clearance space begins to increase. The extra flow available from the pressure component at D will help fill out the increasing space until the flow is equal to the shear flow at F. From F on, the clearance continues to increase, and since there is not sufficient fluid to fill it, the film breaks
Incompressible Lubrication; One-dimensional Bearings 39 down into individual filmlets and continues in that state until fresh lubricant is admitted. The space between the lubricant filmlets is filled with air, vapor, and foam. This is schematically shown in Fig. 3-1 d; Chap. 15 offers experimental evidence of this phenomenon. In partial bearings the situation is similar if the arc ends beyond F. If the trailing edge is before F, the film is complete throughout the bearing. The films are usually complete in thrust bearings. 3. Viscosity. The viscosity of the lubricant in hydrodynamic bearings never remains constant. The viscosity of any fluid varies with both temperature and pressure, and there is also evidence that it varies with the rate of shear. While the variation with pressure is significant only at very high pressures, usually beyond ordinary bearing operation, the u~-U Fig. 3-2. Dynamics of a thrust bearing. dependence on temperature is most pronounced at low and moderate temperatures, the very regions in which bearings operate. When losses are low or temperature levels high, average values may be used. When the conditions listed above are at the other extreme, constant viscosity values may yield unsatisfactory results. 4. Heat Transfer. Not all the energy generated in a bearing is carried away as heat by the lubricant. Part of that energy is dissipated by conduction and radiation via the bearing shell, housing, and journal. No two bearing assemblies are alike in this respect; temperature variation over each of the mating surfaces, the presence of neighboring heat sources and sinks, the complexity of assembly parts, and the effect of windage create a formidable problem, particularly since the hydro- dynamic and the heat transfer problems are interrelated and have to be
40 Theory of Hydrodynamic Lubrication treated simultaneously. Usually, the smaller the bearing and the lower the shear losses the higher the percentage heat lost to the surroundings. 5. End Effects. This subject is still unexplored, but a multitude of experimental data on lubricant flow which refuse to conform to theoretical predictions can be explained only by the effects to be mentioned. These include phenomena such as surface tension at the sides of the bearing, the formation of a meniscus, and the sealing effect that such a meniscus has on the free flow of lubricant out of a bearing. This sealing mecha¬ nism causes the lubricant to flow backward along the sides of the bearing, i.e., in a direction opposite to journal rotation, and to reenter the bearing in the low-pressure region to be recirculated. Another phenomenon is the possible formation of a vena contracta around the annular outlet. These and perhaps other effects make the side leakage usually less than that predicted from theory. The points listed above are all major and general phenomena associ¬ ated with hydrodynamic bearings using incompressible fluids. Some points that assume significance only in certain ranges of operation are: a. Elastic Deformation. Under heavy loading and depending on its structure and assembly, the bearing surface will deform. This in effect will produce a different film shape with a drastic change in bearing performance. b. Turbulence. The Reynolds equation is based on the assumption of laminar flow. High linear speeds, large clearances, and low viscosities will cause turbulence with a resulting rise in power loss, a drop in lubri¬ cant flow, and a shift in the locus of shaft center. c. Thermal Expansion. When bearings undergo appreciable tempera¬ ture changes, when journal and bearing materials have radically different coefficients of thermal expansion, or when journal and bearing are forced to expand against each other, the clearances will not retain their original shape and dimension, and the performance of the bearing will be affected. d. Surface Roughness. When bearings are operated at very low values of hmin, the inherent surface roughness of all materials may have an effect, since the roughness may be of the same order of magnitude as the mini¬ mum clearance. This may not only change the shape of the oil film but carry the operation into a mixed boundary region where the require¬ ments of hydrodynamic lubrication are no more than partially fulfilled. e. Unbalance. Most journals will have some residual unbalance. With unbalance the journal center is not confined to a point but moves along some locus. Thus, the steady state is replaced by dynamic conditions. /. Misalignment. Slight amounts of misalignment are inherent in all journal bearing assemblies. When the degree of misalignment becomes excessive, it is necessary to take this effect into consideration.
Incompressible Lubrication; One-dimensional Bearings 41 Very minor items which do not affect bearing operation to any notice¬ able degree but of which one should be aware are the variation of specific heat with temperature and pressure and the presence of air, foam, and foreign particles in the lubricant. In the solutions of this chapter, none of the points mentioned above is considered. In fact, all these ramifications are minor compared with the radical and from a practical standpoint impossible assumption of a one¬ dimensional bearing, a bearing infinitely long or infinitely short. These solutions, however, are useful for a number of reasons. In the first place, they are given mostly in analytical form with all the inherent advantage over numerical results, of which the bulk of exact solutions consists. Oftentimes they are the only available solutions, a useful guide to how a bearing would possibly perform under certain conditions, and they do provide upper or lower limits. Although the quantitative answers are often at variance with experimental results, they do nevertheless provide a means of studying trends and relationships. ONE-DIMENSIONAL JOURNAL BEARINGS If we assume the bearing to be infinitely long in the axial direction, this implies no variation of pressure in the z direction, or dp/dz = 0. Equa¬ tion (1-12) then becomes (3-1) dx \p dx) dx If the flow due to the pressure gradients in the x direction can be neglected (while retaining the com¬ ponent due to shear)—a situation approached by very narrow bear¬ ings—then the term £C-‘2)-° and Eq. (1-12) becomes dz\p dz ) dx (3-2) It should be noted that this last equation is still a function of two vari¬ ables and that it is a less radical simplification than Eq. (3-1). The above equations are fully defined except for h. While the film thickness in thrust bearings can assume different expressions, its form for an aligned journal bearing is universal. Referring to Fig. 3-3, we have
42 Theory of Hydrodynamic Lubrication OB _ R _ R _ e sin 0 sin (7r — 0) sin 0 sin a P = 0- a = 6- sin-1 sin 0^ OS = ^ sin |^0 — sin-1 ^ sin 0^j = R2 — e2 sin2 0 — e cos 0 h = (S + C) - OB = C + e cos 0 + R — \/#2 — sin2 0 « C + e cos 0 or h = C(1 «+• e cos 0) (3-3) 3-1. Infinitely Long Bearing. The earliest solution of the infinitely long full journal bearing is due to Sommerfeld, who by use of an adroit substitution succeeded in integrating Eq. (3-1). When x is replaced by the angular coordinate 0, that is, when x = Rd, and it is remembered that n is constant, Eq. (3-1) becomes WR% (3-4) de\ de) By integrating once with respect to 0 dp de ~ 6 nURh + Ci h3 II ^3 1^3 0, so Ci II 1^3 6 nURh — ho h3 At some h = ho, Ann h 2i - (3-5) where h is given by Eq. (3-3) and ho is still to be determined. From Eq. (3-5) the pressure is given by _ 6nUR \ f dd h0 f d$ 1 r P C2 Li (1 + e cos 6)1 C J (1 + « cos 0)3J + 2 To integrate the above, let 1 *> 11 n 1 — 1 -f € cos 0 = , 1 — c cos ^ ... . COS \p — € from which cos 0 — 1 — e cos By using sin2 0 + cos2 0 = 1, we have (1 — t2)Yi sin \j/ sin 0 = 1 — c cos \J/ and by differentiating one of the terms above (l-€2)^#
Incompressible Lubrication; One-dimensional Bearings 43 The boundaries 0 = 0 and 0 = transform into the same boundaries in the ^ coordinate, and thus the boundary conditions are p — pa at \f/ = 0 p( 0) = p(2t) (3-6) By evaluating the integrals resulting from the above substitutions, we have / de (i +1 cos ey _ r n -1 cos a» a -€%« J V 1 - €2 J 1 — 6 CO" *#‘ and Thus / d0 € COS ^ d\p = (1 _V)W (* - 2t sin^ + e-J +J8in2^ _ /* /1 — € COS A2 * (1 + € COS 0)2 “ J V l - €2 / (1 - w* ho = n~ 4 sin ®(a) = C2 [ (1 - «2)* C(1 h 0 _ - «2)« V 2e sin \p c2 sin 2^ j 4- C 2 By using the first boundary condition of Eq. (3-6), we have C2= pa By using the second, we get 2C(1 - €2) ho — 2 + €2 (3-7) and thus the expression for the pres¬ sure distribution becomes, by revert¬ ing to the original coordinate, P = Pa + 6pURe (2-f« cos 0) sin 0 C2 (2 + €2)(l + 6COS0)2 (3-8) Fig. 3-4. Pressure distribution in¬ cluding negative regions. where pa is the pressure at 0 = 0. This can be evaluated from the condi¬ tions at B, where a given inlet pressure pi corresponds to a given angle 0i. If the inlet hole is at 0i = 0, then of course pa is the value of inlet pressure. Equation (3-8) yields regions of high negative pressures such as shown in Fig. 3-4. The magnitude of these negative pressures will depend both on the position of 0i and on the magnitude of p\. In any case, the
44 Theory of Hydrodynamic Lubrication pressure distribution resulting from Eq. (3-8) is always antisymmetrical about 0 — t and p = pa. - "• { The vertical load component is/by integrating the pressure over the bearing surface given by W sin <*. = jf" LR de p sin 6 By integration by parts, W sin <f> = LR £ — p cos 6 — j ^ cos 0 dd j and, by using from Eq. (3-5), we have dd cos 6^d$ du . J 6^LUR2[h0 f2' cos 9 de f2 wsm* = —c^[c]0 (i + (cos ey ~ Jo 2r cos 6 dd (1 + C cos 0)2J These can be reduced to the same integrals used in evaluating the pressure distribution by writing cos Q _ 1/c 1 /€ (1 4- c cos 0)z (1 + € cos 0)2 (1 + C cos 0)3 , _ cos 0 _ 1/c 1/c an (1 4~ c cos 0)2 (1 + « cos 0)2 1 4" c cos 0 Thus -«») f2r ^ C2 \ 24- €2 > (1 4- € cos 0)3 d0 : COS 0 | w ^ 6nULR* (2(1 - c2) f2* dd W sm <f> = — j v ' 1 _ f 2(1 — t2) ] [* de [» dfl L 2 + «* J Jo (i + * cos ey Jo i +1< The only new integral to appear is the last one, and its value is de _ i / 1 4- € cos 0 y/\ — e2 By following the same procedure used for evaluating the pressures, we obtain 1V sin 6 - (3 „x w sin <t> ~ (2-|- e2)(i _ t2)H ^ To find the attitude angle, we must look for the load component at right angles to W sin <t>, or W cos <j> = j^'LRde p cos 0
Incompressible Lubrication; One-dimensional Bearings 45 Upon integration by parts, W cos <t> = LR sin 0 — j sin 6 ~ dd j = — LR sin 6 ^ dd By use once again of the Sommerfeld substitutions, ^ GpULR*\ ho ( # c2 . , A . 1 .1 I2' W COS <t> = Ci i) (cos j, - 2- Sill2 + J— COS |o and W cos <f> = 0 or, since W ^ 0, <f> = ^ (3-10) and the displacement of the shaft is always at right angles to W sin <f>. Moreover, since there is no load component at right angles to W sin 4>, W sin <f> = W is the total resultant load. This unrealistic result is a consequence of including the negative pressures in the integration for load capacity. Thus we can rewrite Eq. (3-9) as _pN/R\_ p Vcj " (2 + e2)(l - t2)» . 12rt W'11) and the Sommerfeld number is seen to be a function of c only. The shear stress at the journal is, from Eq. (l-16a), given by — . h dp Tx ~ h + 2R d§ and the frictional force on the journal is de Fj = j*’r,LR de = -LR J*' j- + € COS d M /02'( 1 + ec°s e^de\ By using for the integrals the expressions derived above F - i tt # 4x(l + 2c2) Fj pLL c (2 + €2)(1 _ e2jW The friction factor defined as / = F/W is then f-h-9. I + 2i! (3-13) W ~ R 3f W) At the bearing surface, by Eq. (l-15a),
46 Theory of Hydrodynamic Lubrication The difference in the journal and bearing torques is balanced by the external load IT, which exerts a moment through its eccentricity e or RFj = RFb + We (3-14) The friction in a concentric journal bearing when e = dp/dd = 0 is often referred to as Petroff’s equation and is given simply from Eq. (3-12) by 2 thULR/C, The foregoing analysis yields quantitative results that are far afield from any actual bearing performance. They are also qualitatively in error; for, as mentioned previously, negative pressures of the same order as the positive pressures could not possibly be maintained and the locus of shaft center is never a straight horizon¬ tal line. The major shortcomings of the foregoing analysis can be eliminated by imposing a more realistic boundary condition at the trailing end of the pressure wave. These boundary conditions are discussed in greater detail in Chap. 4. Suffice it to say here that no negative pressures are allowed and the requirement is imposed that, at the point where the pres¬ sure wave falls to zero, line E, the pressure gradient too becomes zero, as shown in Fig. 3-5. Thus the boundary conditions imposed are: vq CTmin Fig. 3-5. Pressure distribution ex eluding negative regions. p = 0 dp _ dd ~ p = 0 = 0 at 6 = 0 at Q = 62 at 6 — d2 (3-15) Since Eq. (3-1) is a second-order differential equation, it cannot in general satisfy more than two boundary conditions. It will be shown, however, that the last two conditions of Eq. (3-15) are a special case of a more general single condition and that the solution based on Eq. (3-15) is only one of a family of possible solutions. Let the last two boundary conditions at 62 be written as the single condition D = P kd9 then, by integrating the Reynolds equation, we obtain QnUR[ f9 dd , „ f9 dd • V = Writing C2 —-7^—3T2 + Ci je> , r.i- + Ci [ (1 + 4 cos 9) -/: * (1 + € cos d)n (i + € cos dy 1_ 1 + € COS 02
Incompressible Lubrication; One-dimensional Bearings 47 the constants C1 and C2 are evaluated by using p = 0 at 0i and P = fcgf atfl2 n h + kg2 n _uglh-9ih tl ~ h + kg* k h + kg* The expressions for p and dp/de then become 6mI/K P C2 f* de _ h + kg2 [• Je, (1 + « cos 0)2 h + kg3 Je, d6 (1 + € cos 6 9*1. dp 1 /2 "h kg d0 C2 [(* + € cos 0)2 ^3 + kg3 (1 + e cos 0)3J Since e < 1, 7n > 0, and g > 0, the denominators never vanish. Thus the conditions for p = 0 and dp/dd = 0 at 6 — 62 are for p = 0 k(g2h - gzI2) =0 or kg2{Iz - gh) = 0 for5? = 0 g(h + kg2) = /3 + kg3 or /3 - fif/2 It follows then that, to have p = 0, we need either k = 0 d0 or /**« d0 /'®I / /i _i_ zr2 = (1 + * cos 02) / Jet (1 + € cos 0)2 J9l (1 + c cos 0)3 Thus if the later condition holds, both p and dp/dd are zero for any value of k including k = 0, which corresponds to our particular boundary conditions. By again employing Sommerfeld’s substitutions and using the first two boundary conditions of Eq. (3-15), we obtain for the pressure distribution 6 nUR I, . (2 + «2)^ — 4e sin \J/ -J- e2 sin $ cos yp ] P = C2(l - e2)* r - ‘ 81,1 * 2[1+«C06(»,-T)1 ) (3-16) where cos ^ = € + cos 0^ an^ ^ corresponds to 02. By using the last J. *i € COS tt condition, namely, p = 0 at ^ = ^2, we obtain from Eq. (3-16) €[sin (^2 — t) cos — x) — ^2] + 2[^2 cos (^2 — x) - sin (*2 - tt)] = 0 (3-17) which determines ^2 and thus 02. Equation (3-16) with ^2 determined by Eq. (3-17) gives a pressure profile satisfying all the conditions of Eq. (3-15).
48 Theory of Hydrodynamic Lubrication For the two load components, by writing ^ = ^2 — a.,,, (1 - «2)(1 + t COS^'j) W sin <b = WW/WW* cos # ~ sin #) . . ►V sin </> (1 _ + { cog (6 19) 3nUL(R/cr r ««(i + cos (1 - «2)*(1 + t cos f'2) [ 1 - e2 + 4(^2 cos i/'i — sin ^2)21 (3-20) 2(1 — <2)^(sin ^2 — *p2 cos ^2) «(1 + cos i.„„ ^ c r vam ^2 Y2 WO Y2J /o m\ tan <f> — .//\2 (3_21) («)/=^± + _^!s_ (322) The use of conditions (3-15) resulted in the elimination of the region of negative pressures and the derivation of a journal locus, Eq. (3-21), which conforms with experimental evidence. Numerical results of Eqs. (3-20) to (3-^2) are given in Table 4-1, where they are tabulated together with the solutions of finite bearings. 3-2. Infinitely Short Bearing. Equation (3-2) has been written down as applying to infinitely narrow bearings. Since for aligned journals h = f(x) alone, we can integrate this equation by treating h as a constant. Thus QnUdh z2 h3 dx 2 _ uflu an z- r r V = TT + C'z + C* By using the boundary conditions p = 0 at +L/2 **•■> - W- (t - *■) ir-r^'w <3-23> This pressure distribution is parabolic in z and antisymmetrical about 6 — nr and p = 0; that is, the region of negative pressures is identical to that of positive pressures. Here the problem of negative pressures is dealt with simply by deleting the region t < 0 < 2tt where the negative pressures occur. (This, of course, can also be done for the infinitely long bearing, and results of such an integration are given in Chaps. 7 and 12.) By summing forces only over the interval 0 < 0 < tt: c sin 0 cos 0 , ad Wx = —2 f f p cos 0 R d0 dz = — [ n , jo Jo 2c2 jo (i +«cos ey o [' fL/2 a T> Ja j nUL* [* € sin2 0 Wy = 2 \ / p sin 0 R d0 dz / y—- -r- d0 Jo Jo 2c2 jo (i + € cos ey By again using the Sommerfeld substitution and integrating,
Incompressible Lubrication; One-dimensional Bearings 49 (3'24a) W‘ = !^W^W> (3-246) The total load capacity is then given by pUL3 e 4C2 (1 - €2) ^ ,t - .M I'**1 - ‘2) + 16‘2]» (3-24c) T (§)’ ’ s (I)’ - .■)' + I6.'jw <«-25> This last expression is seen to be independent of bearing diameter. The effect of diameter is felt through the value of Cy which is usually a function of bearing diameter. The attitude angle is given by tan 0 = ^ ^ ~ (3-26) Since there is no pressure-induced shear, _pU T h and F = J*' M £ LR de = (3-27) with the friction coefficient given by — f — 2t2£ (o og\ CJ “ (1 - €2)^ ( } The lubricant flow out the sides of the bearing is, by Eq. (1-136), f'RVdp Qz~ Jo 12m dz d6 = eULC (3-29) ±L/2 The parameters at the point of maximum pressure are 1 - (1 + 24e2)>* COS 00 = 4e ho = j [5 - (1 + 24*2)»] (3-30) _ _3pUL2 c sin 0O /b oi\ P° 4RC* (1 + t cos fio)3 1 ' This treatment yields a fair approximation to the performance of narrow bearings at low eccentricities and, by the simplicity and compactness of its mathematics, constitutes a useful tool in the analysis of lubrication problems.
50 Theory of Hydrodynamic Lubrication 3-3. Partial Bearings. By definition any bearing having an arc less than 2t is a partial bearing; in practice, however, the criterion is usually an arc less than 180°. The analysis of partial bearings, as evident from Fig. 3-6, is made more difficult by the appearance of two new independent parameters, the load angle a and the arc span 0. Most available solu¬ tions are similar to the Sommerfeld solutions for a full bearing and in¬ clude negative pressures. We shall, however, restrict ourselves to bound¬ ary conditions similar to those of Eq. (3-15), namely: V = Pi dp n v = Te =0 at 0 = 0i at 0 = 02 p > 0 in the region 0i < 0 < 02 p = 0 at 0 > 02 Although the methods and the equa¬ tions involved are similar to those of full bearings, the actual calculations are laborious and the final expres¬ sions are long and cumbersome. For this reason, the results will be given in graphical form. If in Eq. (3-5) the limits of integration are kept general, then P = Pi + 6/tiUR feh- h0 Je, hz or after performing the integrations in the same manner as before V = Pi + rf‘ 6 pUR C2(l -62)* {[(1 <2) c(1 + 2)](* ^ - £e(l - «2) - ^T-°j (sin - sin f,) - (sin 2^ - sin 2^i) (3-32) where \p is the angle from Sommerfeld’s substitution and h0 is to be evaluated from the boundary conditions at 02. Equation (3-32) was evaluated by numerical integration for low eccentricities and by the use of the mathematical expression for high eccentricities. The results are extracts from families of solutions and represent points of maximum load capacity, a condition at which all bearings tend to operate. Figures 3-7 and 3-8 present performance of partial bearings with the load vector located at any arbitrary position with respect to the bearing boundaries
Incompressible Lubrication; One-dimensional Bearings 51 and containing no negative pressures. For a given S and arc length 0 Fig. 3-7 will, if 0 is set equal to 0', determine the optimum load angle a. If 0 > 0', then obviously the load ca- Numbers along curves indicote attitude angle <f> 150° 180° >3 = 210° pacity will be the same, with the arc 0 — 0' only increasing the frictional losses. The condition 0 = 2a rep¬ resents, of course, the case of cen¬ trally loaded bearings. It should 22 r 0 20 40 60 80 100 120 140160 180 200 Load angle a, deg Fig. 3-7. Load capacity of partial bear¬ ings. 0 20 40 60 80100 120140160180 200 Load angle a, deg Fig. 3-8. Friction factor in partial bear¬ ings. be pointed out that here, as throughout the book, the reference area for partial bearings is the projected area of an equivalent full bearing, that is, P = W/LD. 3-4. Fitted Bearings. Fitted or no- clearance bearings are those whose diameter equals the diameter of the journal, as shown in Fig. 3-9. In that case the clearance is zero and Eq. (3-3) becomes h = e cos 0 Fig. 3-9. Notation for fitted bearings. By rotating the 0 axis by 90° in the di¬ rection of rotation, using h = e sin 0 in Eq. (3-4), and expressing the constant ho by e sin 0O, we obtain p(9) = j^sin 0o - In tan - 2 cot 0 j + <7, By using the boundary conditions p(0i) = p(0o) = 0, we obtain for (3-33)
52 Theory of Hydrodynamic Lubrication sin 0O and Ci 2(cot fli — cot 02) Q 1/1 1/ V2J /Q Sin 0 cot 0i/sin 0i — cot 02/sin 02 + ln [tan (02/2)/tan (0i/2)] n 3nU R [8in 00 (s§4l “ln tan I1) "2 cot *i] The integrations for load capacity performed over the entire bearing arc yield the following results: SnULR2 T -Cie2 , . i • n ( n i , 02 Wy — ^ SjHJR ^C0S — cos 2' sm 0 I cos 2 2 — cos 0i ln tan —^ — 2(sin 02 — sin 0i) j (3-35a) Wx = (s^n ^2 sin ^0 + sin 0O ^cot 0i — cot 02 0 0 \ + sin 0i ln tan ^ — sin 02 ln tan + 2 ^cos 0i — cos 02 + ln tan ~ — ln tan ^ j (3-356) j^4 ^ln tan |-2 — ln tan ^ + 3 sin 0o(cot 02 — cot 0i) j e (3-30) Table 3-1 relates the various parameters of a fitted bearing with its performance, under the requirements of maximum load capacity for a given minimum film thickness. This shifts the position of the load vector arbitrarily along the bearing surface. The requirements for Table 3-1. Optimum Conditions in Fitted Bearings 0, deg 30 60 90 120 150 0i 129.1° 92.2° 64.3° 40.7° 19.4° a 17.5° 35.8° 55.7° 77.7° 101.1° a/P 0.58 0.596 0.62 0.65 0.674 hi/hmia 2.18 2.14 2.08 2.0 1.804 c/hm\a 2.18 2.14 2.31 3.02 5.43 h{)/hmia 1.4 1.48 1.62 1.84 2.131 0o 150.1° 136.4° 135.2° 142.4° 156.9° 4> i J 56.6° 38° 30° 28.4° 30.5° w uLNIt* 0.00451 0.01685 0.03309 0.04534 0.03924 _ ^min 0.0407 0.0774 0.1073 0.1284 0.1339 uLNR1 At 9.027 4.597 3.242 2.831 0.3413 ” min
Incompressible Lubrication; One-dimensional Bearings 53 minimum friction in the bearing are very close to those of Table 3-1. It should be noted that in fitted bearings the minimum film thickness is always at the outlet end. 3-5. Floating-ring Bearings. This kind of bearing has a thin ring float¬ ing freely between journal and bear¬ ing as shown in Fig. 3-10. The pur¬ pose of the ring is to reduce the shear losses by decreasing the relative speed between the mating surfaces. Since the losses vary as the square of the speed but vary only linearly with area, the rotation of the ring will always result in some reduction in drag, if all other parameters remain unaffected. It is clear that for equilibrium the sum of the two moments M i and M 2 acting on the floating ring must be zero. The expression for the drag is given by Eq. (3-12), which will be rewritten for our purposes in the following manner: Fig. 3-10. Floating-ring bearing. nULR 4x(l + 2e2) C (2 + «2)(1 — 4tvLRU [ 2 C [(l-€2)“ 2t nULR . (2 + €2)(1 - €2)*J (yjruLUe2 Cy/T=7* ^C(2 + e2)(l -e2)» By multiplying and dividing the last term by (R/C)( 1/e) and recalling the expression for W sin <t> given by Eq. (3-9), we have „ 2TpULR . cC w . * ~ C(1 - t2)» + 2R n * By a similar procedure the drag on a bearing is „ 2rgXJLR eC lir . 6 “ C( 1 - £2)* 2R Sln We thus have for the two moments (3-37) M, = - ^^L(RXN - R2Nr) Cl( 1 - €!2)* 4ir2nRzzLN R ^sin + Wi
54 Theory of Hydrodynamic Lubrication Since R\N — R2Nr « R2(N — Nr) and by writing W = 2LR2P and € = cos 6, we have by equating Mi to M2 „ _ . 1 - <.2)» v ^ - 1 + C2fl23(l - ‘22)k + P[C2t2(1 “ <s2)W + Cltl(1 - It is shown in Chap. 1 that the behavior of a unidirectionally loaded bearing in which both journal and bearing rotate in the same direction is the same as though the journal alone were to rotate at a speed equal to the sum of the journal and bearing speeds. Thus, according to Eq. (3-11) _ fRi\*n(NR + N) _ (2 + €l’).(l - €1*)K P 12t26i s2 = -(2+€22)(i -ef,)* 12t262 By using these relations to eliminate P in the expression for N, we have for the speed of the ring and for the relation between the two eccentricities - 3*i [e,(l - c,*)» + £-2 «2(1 - <2*)» j (3-39) The total power loss for a floating-ring bearing is " - [m; w”n*• *cfift)*<-u-v*\u <3-"» Figure 3-11 shows graphically the relation between the various ratios of clearance, radius, and speed. If one considers an equivalent standard bearing, denoted by subscript e, as having the dimensions Ne = N (\ = Cl Re = Ri = R then by using the more critical outer film as a basis for comparison, it results that the floating ring has a lower load capacity if Se/S2 > 1 and a higher capacity if Se/S2 < 1. The power losses are always less in the floating-ring bearing. 3-6. Porous Bearings. In a bearing made of a porous material, lubri¬ cant flows out of the bearing surface with a certain velocity VQ. If
Incompressible Lubrication; One-dimensional Bearings qy is the rate of flow per unit area, then oo dp q«=-Ty - = V0 v-o M where dp/dy is the pressure gradient at the bearing surface and $ is a property called permeability which varies with porosity and size of pores. Its dimensions are square inches. From the requirements of continuity we have for the porous matrix Vq = * V‘p = 0 so that., since ^ 0, ”o 0.2 0.4 06 0.8 1.0 1.2 1.4 1.6 1.8 2.0 y2p = q Rotio °f cleoronces, Ct/Ct The problem then is to solve Reyn- J'10-.3'1Var‘atiof.of rin« 8Peed with r , . bearing characteristics, olds equation for the pressures m the oil film simultaneously with that of Laplace for the porous matrix with a common dp/dy at the boundary, or *(»%)+* (V^\ = wdh 2*0j> I (3-41) dx\v dx/ dz\u dz) dx m dy |y=o d2p d2p d*p _ 3x2 ^ dy2 dz2 Two assumptions are made in solving this set of equations: 1. The bearing is infinitely short. 2. dp/dy is linear across the matrix and is zero at the outer surface of the porous bearing shell. Assumption 1 yields d2p/dx2 = 0; assumption 2 gives d2p dy2 Hence from Laplace’s equation = const = K d2p _ dz2 = -K and dp dy KH = -£?l dz2 H y = 0 OZ |y-0 where H is the wall thickness of the porous bearing. By using this expression in Eq. (3-41) with ^ ^ = 0, we have (A. + or d2p _ dz2 6m U e sin Q RC2 (1 + € cos ey + 12$>H/C r-TTii (8-42)
56 Theory of Hydrodynamic Lubrication Equation (3-42) solved and integrated for the resultant load along the line of centers yields W§ =M^V11±_* 2C2 ^ 6A.2 111 [fc2 - fc(l fc2 - k( 1 + €) + (1 + «2)] *) + (!- <)2 (1 + *: + «)* J (! + *-«) + (tan- - ‘I - tan- P-^ - -UP th? y/'& ( L fcV3 V3j L *:V3 vUJ/ (3-43) where k* = V2’t>H/C3. Figure 3-12 gives a comparison of the relative load capacities of porous Fig. 3-12. Load capacity of porous Fig. 3-13. Friction factor for porous bearings. bearings. and solid bearings both based on the short-bearing approximation. The coefficient of friction as given by -/ = n j 2ir2<S 1 + o sin CJ ' H -s/l - e ' 2 is plotted in Fig. 3-13 as a function of il, the relative load capacity of the porous and solid bearings. ONE-DIMENSIONAL THRUST BEARINGS The analysis of thrust bearings is made somewhat easier by the simplic¬ ity of the expressions for film thickness and by the less complicated bound¬ ary conditions. In most thrust bearings the film is nondiverging and continuous and the problem of negative pressures does not arise. The
Incompressible Lubrication; One-dimensional Bearings 57 pressures at both the inlet and outlet edges are simple boundary values, usually atmospheric. In fact, it can be shown that the elaborate con¬ dition p = dp/dx = 0 at the outlet cannot be satisfied for bearings having a converging film shape. A number of these one-dimensional thrust bearing solutions were obtained by Lord Rayleigh, who also derived parameters for optimum film shape and bearing proportion. The quantitative results here are no closer to reality than in the case of journal bearings, but, again, many of the dimensionless groups and much of the qualitative behavior deduced from one-dimensional analyses are also valid on the basis of more exact solutions. In reality a thrust bearing, which is designed to support axial instead of radial loads, looks like the device of Fig. 3-14. The popular simplifica¬ tion is to treat the sectorial segments of which such bearings consist as a simple plane slider as shown in Fig. 3-2. By using the system of coordi¬ nates given in Fig. 3-2, where U is negative with respect to x, Eq. (3-1) reads 6S| Fig. 3-14. Thrust bearing. <£ = -wh- ax — ho h3 From this with p(hi) = p(h2) = 0 Integration of the expression for load capacity by parts yields W-L^pto-L („ i;; - /; o, w-toVL(J*’-£ The center of pressure 1 can be obtained from The expression for shear stress is, from Eq. (l-16a), nU , h dp h 2 dx (3-44) (3-45) (3-46) ■dx (3-47)
58 Theory of Hydrodynamic Lubrication and by using the expression for dp/dx from above, we have , 4h — Sho h2 '■-'“(‘/Tt -“•£'!) (3-48) By using these general equations, a number of solutions for bearings having different film shapes can be obtained. 3-7. Plane Sliders. This bearing configuration is shown in Fig. 3-2, and its film thickness is given by , h2{a — 1) h = ax = ——~ x Jt> (3-49) where a = h\/h2. By integrating between X\ = hi/a and x2 = h2/af 0.20 0.18 Cf .0.16 Lp 0.14 0.12 0.10 /' 1.0 0.9 0.8. Cf 0.7 0.6 .0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 a - hjht Fig. 3-15. Load capacity and friction in plane sliders. we have for the pressure distribution V = _ 6fiU (hi — ax)(ax — h2) (hi + h2)x2 h-l > x> -2 and for the parameters of maximum pressure t 2hih2 2d t 0 ” hi + h2 ~ T+~a k2 3nUa(a - l)2 2aa Integration of Eqs. (3-46) to (3-48) yields V o = ^ - tt1’ (rh)' ['” * - 2-r^r] - ~cPc' (^)B(a2 — 1 — 2a ln a) (a2 — 1) ln a — 2(a — l)2 = 2(a2 - 1) ln a - 3(a - l)2 ixUBL (3-50) (3-51) (3-52) (3-53) (3-54) (3-55) B 3(a + 1) ln a — 6(a — 1) h2 The dimensionless coefficients Cp and Cf as a function of a are plotted in Fig. 3-15.
Incompressible Lubrication; One-dimensional Bearings 59 By setting in Eq. (3-53) dW/da = 0, the value of a for maximum load capacity is obtained, and its value is5 a = 2.2. This is also evidenced from Fig. 3-15. Based on a = 2.2, the values of the other quantities are W = 0.1602 ho = 1.37hi F = 4.7 ^ h>2 Jo If the bearing is made up of a series of tapered-land segments, each segment being equal to B/n, while the total breadth is B, then which means that the bearing load capacity is (1/n) times smaller than that of an equivalent unbroken surface. This, of course, assumes that the taper over each segment B/n is equal to a. 3-8. Curved Sliders. If h = mxni then the value for the resultant load is _ 6pULB* I"(3n - 1 )(a-2+1'" - 1 )(a-3+2'" - 1) a~2+2'" - I] V(a1/n - l)2 L (2n - l)(a“*+l'")(3n - 2) 2(n - 1) J (3-57) By setting n = 2, that is, by using a parabolic surface and calculating its load capacity for the optimum value of a, which is 2.3, we have W = 0.163 If the film thickness is expressed exponentially, h = e?x, then by perform¬ ing the rather easy integration, we obtain t\r _ 3uUL Ta2 — 1 /3(a2 — a3)B~\ (0ah2)2 [ 6 ■*" a3 - 1 J or since = ln a, <-> For an a = 2.3, W assumes a maximum value of W = 0.165 h 2 The similarity between the load capacity for all these various bearings underlines the important fact that, once hi and h2 are fixed, the exact shape of the oil film is not of great importance. This conclusion is reached on the basis of two-dimensional analysis also.
60 Theory of Hydrodynamic Lubrication 3-9. Step Bearings. If in Eq. (3-1) we set h = const and consider isothermal conditions, the resultant pressures are zero; thus a bearing having a constant film thickness has no load capacity. However, if the film is parallel but has a step in it such as shown in Fig. 3-16, the bearing will develop hydrodynamic forces. The expressions for the film thick¬ ness here are By evaluating this expression for region Bi with the coordinate system Fig. 3-16. Step slider, as shown in Fig. 3-16 we have p = 0 for x — 0 V — Vo for x = Bi where pe is the common pressure at the step. The two conditions yield Similarly, for region B2 with the coordinates shifted to the leading edge, we have h = hi in region Bi h = h2 in region B2 By integrating Eq. (3-1) for a constant h, we obtain p = &pU^^x + Ci B Ci = 0 pc = 6pU^~^Bi C» = 0 Since pc is the same in both cases, we have or (3-59)
Incompressible Lubrication; One-dimensional Bearings 61 and the pressure profile is h2(Bih2* -j- B2hi*) P(I) " Ti?- / \ 6pU P(I) - T? 1 B,A,» + S2/n8 hi{Bih2* -f- B2hi*) B1A2* -I- B2hi* - 1 x for region £1 (3-60a) x for region £2 (3-606) which reduces to zero if hi — h2 or if B1 or £2 = 0. It is apparent from Eq. (3-1) by setting h equal to hi or h2 that the pressure gradient is dis¬ continuous at the step. The load capacity is „ = wt + w. = [*'xdx + W(h>-h2)L f*>xdx hi* Jo h2* Jo w 3pULBiB2B(hi — h2) ( . w B557(3'61) By writing £1 = £ — £2 and a = hi/h2j Eq. (3-61) can be rewritten _ 3pULBB2(B - B2)(a - 1) (Bja* + B — B2)h2i To find the optimum B2 and a we set dW n , dW n aF2 = 0 and 3a The first relation yields dW _ 3pULB(a — 1) dB2 W [*(£20* + B - £2)(£ - 2£2) - £2(£ - £2)(a3 - I)] _ n L (£2^ + £- £2)3 J " U and, to satisfy this, £2 - 2££2 + B22 _ /£ — £2V V ) £22 or £ = £2(a* + 1) The second relation yields dIF _ 3pULBB2(B - £2) I"£2a3 + £ - £2 - (a - l)3£2a2] da W [ (£*a* + £ - £2)2 J which yields £ = B 2 2a3 - 3a2 + 1
62 Theory of Hydrodynamic Lubrication By using the expression for B from above, B2 = £2(a* + 1) or 2a3 - 3a2 + 1 (a - l)(4a2 - 8a + 1) = 0 One root is a = 1 and the other one is a = 1 + V% = 1-866 The optimum value of B\/Bi is then §! = o2(2a - 3) = 2.549 r>2 With these optimum parameters, the performance of the step bearing is w 0.2052Mt/L£2 h2* B = 0.4262 / = 4.091 ^ It can be shown by the use of calcu¬ lus of variation that a stepped film is the optimum film shape for a slider bearing. This is true also for the more general case when the viscosity of the lubricant is considered to be a function of pressure. 3-10. Composite Bearings. The bearing termed “composite” is made up of a combination of tapered- and flat-land bearings, as shown in Fig. 3-17. Its film function is h = ax for region B\ Fig. 3-17. Composite slider. h = h2 for region B2 For region B2i taking the exit edge as the origin, we have for boundary conditions p = 0 at x = 0 p = pc at x = B2 with the common pressure pe given by Pc = 6/i t/f? 2 For Bi, writing a in terms of the taper hi — h2f > hi — h2
Incompressible Lubrication; One-dimensional Bearings 63 and by rewriting the expression p = f(x) into p = }{h) dp _ dp dh _ h\ — h2 dp dx dhdx Bi dh we have dp_^UBl(ha 1\ dh hi_htyh3 h,j and upon integration „ru\ SpUBifl h0\ n rW = h^h2{h-2h*) + C2 The boundary conditions for this equation are: p = p< = 6hUB2 at A = A2 hi p = 0 at h = hi By using these two conditions, we obtain p = 6»U |b2 (^-t - ^ ^ - 2^-^ j J (3-62) i. _ 2hih2(Bihi + B2hi) 0 (A, + A2)[B,A2 + 2B2A,V(Ai + A*)] ( ) It should be noted that the expression for pc is a function of the local h alone, and thus p will be continuous at the boundary. The total load is given by ■dh ri2 fB* fhl B i W = / Lpdx — / Lp yo fti — a2 _ 6Ml/LB!2 (B2 /B2 \ /A + 1 1 A + 1\ A2* (B,VBi /Vfc + 2 2 7* + 2/ -rh + i[^-m]| <"*> where k = ^-=A2 7 = r>7 ^/B')kik + D 2(Bi/Bi)(k+ 1)2+ (k + 2) With, --.</($-1) [(& - k) {; •"+j; (3 w - c) H l'“i2i:[3<,-’>rT5-2]-<h' _,[M- A: + 2 + 1) 6_ A* —|— 2 (3-65)
64 Theory of Hydrodynamic Lubrication If in the equation for ho we set B2 equal to zero, the expression reduces to that of an ordinary plane slider. With B2 > 0 the value of ho is less than for tapered-land bearings, which means that the point of maximum pressure has shifted closer to the exit edge. However, since ho > h2 this peak will always remain in the tapered portion of the bearing. The load capacity of a composite bearing is at an optimum when Bi/B2 = 5, at which point it will be 25 per cent higher than for an equivalent ta¬ pered-land bearing. Its losses, how¬ ever, will always be higher, as can be easily deduced from the longer span of its minimum film thickness h2. 3-11. Pivoted-shoe Bearings. The pivoted-shoe bearing is one which, instead of being rigidly fixed with respect to the runner, is sup¬ ported by a pivot and is thus free to assume any inclination. This it will do in a manner that will yield the highest load capacity. The per¬ formance of these bearings can be calculated from the results obtained for plane sliders. I n a pivoted-shoe bearing the resultant force obvi¬ ously has to act through the pivot, and this pivot position must coin¬ cide with the center of pressure £ derived for the tapered-land bear¬ ing. Each £ is associated with a certain minimum film thickness h2, and h2 in turn is related to load ca¬ pacity. Thus, from Eq. (3-54) [6) Pivot locotion Fig. 3-18. Performance of pivoted sliders. (H)#(a2 - 1 - 2a ln a) (a2 - 1) ln a - 2(a - l)2 For a given pivot position, a can be found from Eq. (3-54) and W from Eq. (3-53). Since these relationships are implicit in a, the results for W = /(a,£) are as given in graphical form in Fig. 3-18. The pressure distribution in plane sliders is always asymmetrical with respect to B/2. The resultant force will thus always be off center, and it follows that in order to have any load capacity, the pivot must be located at £ > B/2 from the inlet edge. It remains, however, a fact of
Incompressible Lubrication; One-dimensional Bearings 65 pivoted-bearing operation, both journal and thrust, that the bearing will operate satisfactorily with the pivot located in the center. Various explanations for this behavior may be offered, and some of the explana¬ tions are discussed later in the text. One of the more obvious reasons is that the surface of the pads is never flat both because of geometrical imperfections and because of the cantilever effect, which tends to bend the pad surface into a circular or parabolic shape. Fig. 3-19. Curved pivoted slider. In the following analysis it will be seen that a value of £/£ = % yields load capacities if the pads have some curvature. The equations are derived for a parabolic pad surface, but it can be shown that, since the change in elevation, hc of Fig. 3-19, is very small, a parabolic and a circular arc will yield similar values of hc as a function of x. By putting the coordinate axis at the leading edge, we can write he = hc — 4//c(* - y2y and for the film thickness h h = hi — x(hi — h2) — hc = h2{b[4:(x — %)2 — 1] + (a — ax + x)) where £ = x/B, b = He/h2, and a = hi/h2. By substituting h in Eq. (3-1), we get where the constants of integration can be evaluated from p(0) = p(l) = 0. The load capacity is given by
66 Theory of Hydrodynamic Lubrication W = B P pLdx = m^2L Ci. yo /i 2 (3-66) where Cp is a result of numerical integration and is given in Fig. 3-20. The load in Eq. (3-66) is given in terms of the outlet film h2 which, 0.66 Fig. 3-20. Load capacity of curved pivoted sliders. (a) Load coefficient based on minimum film thickness; (b) load coefficient based on outlet film thickness. Fig. 3-21. Pivot position for curved sliders. however, is usually not the point of minimum film thickness. The latter can be found from ^ = 8x6 - 46 - a + 1 =0 dx which yields for the minimum film thickness the following: - ®_-j m1 h ~h rfL+j _ - d2 _ j 8b + 2 2 [ 2 1C6 J »UB2L[a+\ (a- l)2 .I’ nUB*L _ [-2 m~~b\ Cp = -hi~Cpi Thus W = hi Cpi too is given in Fig. 3-20. The coefficients of the remaining parameters £ p nUBLn B Fi ~ -J— Cfl hm\n C f 1 f _ ,vmin i 3 ■ ~B C7i are all given in Figs. 3-21 and 3-22.
Incompressible Lubrication; One-dimensional Bearings 67 This analysis is an example of thrust bearings where negative pressures do occur. From hmin on, the clearance diverges and, depending on the value of He and other parameters, larger or smaller regions of negative pressure may appear. The integrations above did not exclude them, and thus the combination of pad curvature and negative pressures com¬ bined to give pivot positions on either side of the pad center. Over the range of f/B from 0.5 to 0.6 the optimum value for Hc/hmia is about 0.35. For this value of He/hmin the optimum pivot position is f/B = 0.55, and this yields a 10 per cent increase in load capacity as compared to an optimum flat-land pivoted bearing. Fig. 3-22. Friction in curved pivoted sliders. An additional advantage of this type of a bearing is the ease of starting; for even at rest a converging wedge exists between runner and leading half of the pad. Thus upon inception of motion, hydrodynamic pressures build up much faster than between flat surfaces. SOURCES 1. Sommerfeld, A.: Zur hydrodynamischen Theorie der Schmiermittelreibung, Z. Math. u. Physik, vol. 60, p. 97, 1904. 2. Sedney, R., A. Chames, and E. Saibel: The Reynolds Lubrication Equation with Smooth Outflow, Proc. First Natl. Congr. Appl. Mech. 3. Cameron, A., and Mrs. L. Wood: The Full Journal Bearing, Proc. Inst. Mech. Engrs. (London), vol. 161, p. 59, 1949. 4. Du Bo is, G. B., and F. W. Ocvirk: Analytical Derivation and Experimental Evaluation of Short Bearing Approximation for Full Journal Bearings, NACA Rept. 1157, 1953. 5. Lee, J. C.: Analysis of Partial Journal Bearings under Steady Load, ASME Paper 55-LUB-l, October, 1955. 6. Kingsbury, A.: Optimum Conditions in Journal Bearings, ASME Paper RP-54-7, 1931. 7. Morgan, T. V., and A. Cameron: Mechanism of Lubrication in Porous Metal Bearings, Conf. on Lubrication and Wear, Paper 89, London, 1957. 8. Rayleigh, Lord: Notes on the Theory of Lubrication, Phil. Mag. vol. 35, no. 1, 1918. 9. Abramovitz, S.: Theory for a Slider Bearing with a Convex Pad Surface; Side Flow Neglected, J. Franklin Inst., vol. 259, no. 3, 1955. 10. Shaw, M. C., and E. F. Macks: “Analysis and Lubrication of Bearings,” McGraw-Hill Book Company, Inc., New York, 1949.
CHAPTER 4 INCOMPRESSIBLE LUBRICATION; FINITE BEARINGS The various analyses of journal and thrust bearings presented in the preceding chapter are applicable to either infinitely long or infinitely short bearings and are based on the Reynolds equation with one of its pressure derivatives equated to zero. The value of these solutions lies primarily in the fact that they are given in analytical form, that they indicate trends, and that they often establish upper or lower limits to bearing performance. However, for an accurate and correct representation of the finite bearing, we must return to Chap. 1 and start with the full Reynolds equation as given by its various forms (1-6) to (1-12). The present chapter is intended to treat the full two-dimensional Reynolds equation. Owing to the difficulty of obtaining analytical solutions, a substantial portion of this chapter will be devoted to numerical methods applicable to the solution of bearing problems. The results given in the various tables are based primarily on these numerical solutions. FINITE JOURNAL BEARINGS The difficulty in obtaining satisfactory solutions of journal bearings lies not only in solving but also in defining adequately the boundary condi¬ tions of Eq. (1-12). The lubricant can be admitted to a bearing at any angle, and the deeper this angle lies in the converging film the more pronounced is its effect on the resulting pressure distribution. The lubricant is not always admitted at atmospheric, or zero, pressure but may have any value. The situation at the trailing end of the pressure profile is illustrated in Fig. 3-1. In finite bearings this is further compli¬ cated by the fact that the points C, Ef F, and G do not, generally, lie on a straight line across the bearing. These points form curves that are con¬ vex in the direction of motion, and the pressure profiles near the bearing sides may not have the negative pressure loop at all. Also, line B, depending on the mode of lubrication, may not be a straight line. The manner in which the trailing boundary conditions are treated in the literature falls into three general categories. 68
Incompressible Lubrication; Finite Bearings 69 1. p = 0 at the physical end of the surface.* For a full bearing this will be at 0 = 0i + 2x and for a partial bearing at 0 — 0*. This condi¬ tion, curve 1 of Fig. 4-1, yields negative pressures over extensive portions of the bearing. 2. Same conditions as (1), but the negative pressure region is then dis¬ carded and p = 0 is set wherever the analysis yielded negative values (curve 2). If the leading boundary condition in a full bearing is p = 0 at 0 — 0, this approach yields p — 0 over the range x < 0 < 2x. 3. p = 0 for all 0 > 02 and dp/dd = 0 at 6 = 02 (curve 3). This condition requires that the pressure and pressure gradient be zero at 02. All of the above conditions disregard the negative dip EFG. This simpli¬ fication is permissible, for the negative pressure loop is quite insignificant. Boundary condition 1 yields results that are at variance with all available experience, and any calculations based on this assumption are at best of only academic interest. Condition 2 is more realistic, although quantitatively still in error. The* retained positive pressures are substantially reduced both'in angular extent and in magnitude as compared to the pressure distribution obtained by the exclusion of negative values. Condition 3 yields results that compare favorably with practice. Analytical Methods The following general treatment will point out some of the possibilities and complexities in dealing with the Reynolds equation. The nonhomo- geneous, linear differential equation can be easily transformed into dimen¬ sionless form with the inverse of the local Sommerfeld number replacing the dependent variable p. It can be modified to give greater accu¬ racy around the region of maximum pressure where the gradients are steep. In general the approach to obtaining a solution is to assume the existence of a homogeneous and a particular solution, the latter being the solution of the infinitely long bearing. The homogeneous solution, if obtained, would represent the correction factor to the infinite case. Although the homogeneous equation can in turn be expressed as a product * Throughout the book, whenever we say zero pressure, we mean ambient or atmospheric pressure. Fig. 4-1. Boundary conditions in jour¬ nal bearings.
70 Theory of Hydrodynamic Lubrication of two functions of d alone and z alone, it is this homogeneous equation that presents the stumbling block to analytical solutions. Let us in Eq. (1-12) use the substitutions - 2 h p(C/R)2 m x = Re 2 = 779 h = j; V = 0 = =1 L/2 C ixu) then we obtain £(*«)+(?)■*(*■«)-S <«> Since the variation of the pressure gradients is most pronounced around p0, greater accuracy will be obtained if the curves are flattened in that neighborhood. Remembering that h is at a minimum near p0, this can be accomplished by writing p — hap For an aligned shaft, h = f(d) alone; thus d*p, de2 "r /D\*d*p ,3-2 adhdp a |\0 s/dftV , r d’fil . (l) d + -(-+wjp The first derivative of p can be eliminated by choosing a = % and remembering that d2h/dd2 = 1 — h £2 /cv de2 VL/ g + a(9)p = 6(9) with a(0) and b(d) corresponding to the appropriate terms in Eq. (4-2). The original boundary condition of p = 0 at the start of the pressure wave and at the sides of the bearing also holds for p. Since p = h^p h 7* 0 p disappears simultaneously with p. From the expression de 2 dep + de it follows that the condition p = dp/dd = 0 also implies dp/dd = 0. Thus, the boundary conditions are identical for both p and p. The solution of Eq. (4-2) can be assumed to have the form = Q(e>2) + pW (4-3)
Incompressible Lubrication; Finite Bearings 71 where p(6) is the solution of the infinitely long bearing and q(d,z) is the solution of the homogeneous part of Eq. (4-2) d*q 3$2 ‘r Equation (4-4) can be carried further by representing q(d,z) as a product of two single-variable functions q(e,z) = e(e)Z(z) (4-5) By using this in Eq. (4-4) and transposing the variables, we have where X is a constant. This yields two differential equations of single variables 0" - (A [(§)* + 2h(l -«)]+*} © = 0 (4-6a) z"+ (b)2 xz = 0 (4’66) These last equations are not readily solvable. It should also be remem¬ bered that the pressure distribution is not the final goal but that the function p(0,z) and its derivatives must also be integrable before it can yield useful expressions for load capacity, flow, and bearing friction. In the following analyses we shall apply the above outlined general procedure to the solution of two specific bearing problems with two differ¬ ent boundary conditions, that of an axially and that of a circumferentially lubricated journal bearing. Like the Sommerfeld solution in Chap. 3, these too are based on a full and continuous fluid film. 4-1. Journal Bearing with Axial Feeding. For analytical purposes there is no need to nondimensionalize the Reynolds equation, and we can proceed with the form given by Eq. (1-12). This equation with h expressed in terms of 0 gives ^[(l + «C06*)'g]+^[(l+.cw*)'fjf] = -^‘sinO (4-7) where z = z/R. The boundary conditions corresponding to those used in the Sommer¬ feld solution for the infinitely long bearing are ap(-x,2) _ dp(r,S) / ^4'8^ de 36
72 Theory of Hydrodynamic Lubrication For the solution to Eq. (4-7) we write the form given by Eq. (4-3), namely, p(0,z) = q(6, z) + p(0) where p(0) is the particular solution to Eq. (4-7) and represents the solu¬ tion to the infinitely long bearing as given by . v _ 6pURt (2 + € cos 0) sin 0 "" C2(2 + c2) (1 + € cos 0)2 and q{0}z) is the solution to the homogeneous equation h [(1+*cos •>* 3]+1 [(1+£ cos fl)* 8] =0 (4-9) The boundary conditions for Eq. (4-9) are q(—r,l) = q(r,2) q(e,~ = -p(0) 8(-x,*) _ g(,t+L\_pW ae ae Now by writing q(0tz) = 0(0)Z(z) we obtain by substitution into Eq. (4-9) the following two differential equations: d2e 3c sin 0 d0 , A {A lnN — 1 i n + XO = 0 (4-10) d02 1 + c cos 0 d0 g-XZ = 0 (4-11) From Sturm-Liouville theory it can be shown that Eq. (4-10) with the given boundary conditions has a set of real nonnegative eigenvalues X,. The solution to Eq. (4-11) is then Zi = Ai cosh \/\{ z H- Bi sinh \/xt z X, ^ 0 Z o = A0 + Bqz \i = 0 The solution to Eq. (4-10) can be found by using the substitution 0 u = sin2 - which transforms Eq. (4-10) into -!)(«- + [4«* - (4 + l)u + - X (u - 6 = 0 (4-12a)
Incompressible Lubrication; Finite Bearings 73 This equation is of the form of Heun's equation which in standard nomen¬ clature reads z(z — 1 )(z — a) -f {(at + 0 + l)^2 — [at+ 04-1 — 5 + a(y + 5)]z + or) ^ - o)y = o (4-126) where a, g, a, 0, 7, and 5 are arbitrary constants. If the singularities at 0, 1, a, and «> are regular, and with exponents 0 and 1 — 7 at z = 0, the two solutions to Eq. (4-126) for noninteger 7 are y = F(a, g; a, 0, y, i; z) y = zl->F{a, gt; a — y + 1, 0 — y + 1, 2 — y, S; z) with _ a&9 + (1 “ 7)[(« + 0) + b(a — 1) 4- (1 — 7)] 91 (« - 7 + l)tf -7+1) where the Heun function is given by F(a, g; a, 0, y, S; z) = 1 + £ C,z" 1 with the following recurrence relations 07C1 = 0(00 2a*y(7 + l)Cs = (aft/)2 + [(« + 0 4~ 1 “ b) + (7 4“ 5)a]at00 — ayat0 a(n + l)(y -h w)C,h-i = [(a + 0 + n — 6)n + an(7 + S + n — 1) + «0p]C'n — [(n- — l)(a + 0 + n — 1) + a0]Cn-i It can be shown that the Heun function is convergent for all \z\ < \a\ if |a| > 1. Now by comparing Eq. (4-12a) with the standard form of Heun’s equation, we have a = ^ 9 = L5T a> 0 = ± 0 + 4X)*] t-H *-H and since e < 1, we have a > 1 and the Heun functions are convergent for all $. The general solution of the homogeneous equation then by analogy with the Heun relations is so q(fi,S) = QoZo + ^ (A.,- cosh \A< z + sinh \/Xi z)(CiFi + ZW<) (4-13)
74 Theory of Hydrodynamic Lubrication where p „ Tl + < 1 + *. 3 + (9 + 4X,)M 3 - (9 + 4X,)*4 1 1. «1 Fi = F [-IT ~2T’ 2 ’ 2 2’ 2’ Sln 2J * -(*I)'[&«>| J;*■ I] Z0 = A o “h 0o = Co “I- Do j (1 -|- c cos 0)“3 dd = Co + 2>. |(^^i tan- tan |] € sin 6 3c sin in 0 J L + c cos 0) J 2(1 - c2)(l + € cos ey 2(1 - c2)2(l and Aij Bif C„ and D» are constants. Since q(d, z) is an even function in z, we have Bo = Bi = 0 Since p(0) is an odd function of 0, q(0,z) will also by the last two boundary conditions be an odd function of 0. In addition, F, and Hi are respectively even and odd functions of 0. This implies Co = Ci = 0 From the first two boundary conditions we have Do /; (1 + c cos 0)”3 dd = 0 and DiHi(d) =0 and since the integral is different from zero and also D, ^ 0, we have D0 = 0 and Hi{ir) = 0 The solution thus becomes after absorbing D, into At- 00 p(d,z) = p(d) + ^ Ai cosh y/\izHi(\i\6) (4-14) » = i where X, are the roots of * ri + c _ 4 + (9 + 4X)* 4 - (9 + 4X)W 3 1,1 „ F 019 2—' 2 ’ 2’ 2 J = The remaining constant Ai can be evaluated by noting that i/,(0) is orthogonal with respect to (1 + c cos 0)3 over the interval (—that is (1 + € cos d)3Hi(6)Hj(d) dd = 0 for i j* j
Incompressible Lubrication; Finite Bearings 75 By multiplying Eq. (4-14) at z = L/D by (1 + c cos 0)3i/,(0) and inte¬ grating over (—7r,ir), we get [* (1 -f c cos 0)3p(0)//,(0) dd = J -*• cosh [a/>w (L/D)} (1 + € cos e)3Hi\e) de Equation (4-14) is an expression for the pressure distribution in full journal bearings. Like its equivalent one-dimensional solution p(0), it (1) includes negative pressures, (2) is antisymmetrical about 6 — tt, and (3), if integrated over 2v for load capacity, would yield a shaft locus at right angles to the resultant load. 4-2. Journal Bearing with Circumferential Feeding. The differential equation used for the preceding analysis applies here also. With z replaced by its dimensional quantities, Eq. (4-7) reads A[(i + tcoe*)*g] + **![(l + .co8 9)* |f] = (4-15) For a journal bearing lubricated through a circumferential groove the boundary conditions are p(0,z) = p(2ir,z) p = 0 (4-16) t (0,z) = Ye (2^) P (*,-§) = P. where p, is the supply pressure of the lubricant. We assume the solution to consist of three parts: p(0), that is, the solution to an infinitely long bearing; q(6,z), the solution to the homo¬ geneous equation; and p(z), which is due to the supply pressure p,. This last term is usually small and can in most cases be neglected. Thus for the solution of Eq. (4-15) we have . v 6pURe(2 + e cos 0) sin 0 V . M . , (\ z\ p{e’z) = C»(2+ <*)(!+< cos 0)* " 2/ 4"(2) Sln m9 + P‘ (2 " l) m — 1 (4-17) where the form of q(B.z) is assumed to be given by the second right-hand term of Eq. (4-17). The first and third terms of the right-hand side of Eq. (4-17) satisfy the Reynolds equation. The second term can be made to satisfy Eq. (4-15) by inserting it into the differential equation and equating all the coefficients in sin md to zero. From this equation the following recur-
76 Theory of Hydrodynamic Lubrication rence formula for Am(z) is obtained: 2(Z)2 - l)Ai(z) + «(D* + 2) A,(2) = 0 (4-18a) and in general 2(Z>2 — m2)Am(z) + €(D2 — m2 — m + 2)Am_i(z) + €(Z>2 - m2 + m + 2)Am+l(z) = 0 (4-186) where D2 = ft2 ^-5 dz2 We shall assume that in Eq. (4-18a) A 2(2) is related to A 1(2) by a constant — K\, that is A 2(2) = -KxAi(z) with Ki a positive constant to be determined. We thus have for Eq. (4-18a) [(2 - K*)D* - 2(1 + K^A^z) =0 or (D2 - K^A^z) = 0 ,hereK,.-2a±|j) The solution to the differential equations above is straightforward and consists of terms in cosh (K&/R) and sinh (K&/R). Since the homo¬ geneous solution must be even in z, we discard the sinh (K&/R) term and write Ai{z) = Ci cosh ^ (4-19) and also A2(z) = — KiCi cosh The A2(z), At(z), . . . , Am+1 terms are thus given by 2(2)2 _ 4)^(2) + t(£>* _ 4)A1(z) + tD2A2(z) = 0 2(2)2 _ 9)4,(z) + e(2)2 - 10)A2(z) -f «(2)2 - 4)A,(z) = 0 (2)2 - m2 + m + 2)Am+l(z) = - j (D* - m*)Am(z) — (D‘ — m2 — m + 2)Am.i(z) (4-20) Thus, for example, A3(z) and -44(2) are given by - (“' - >) (¥) * cosh ** + C,
Incompressible Lubrication; Finite Bearings 77 Equation (4-20) is a second-order differential equation for Am+i(z), and there will always be two new arbitrary constants to satisfy the boundary conditions. We shall now apply the boundary conditions of equation (4-16). First, it will be noticed that Eq. (4-17) by its formulation satisfies the first and second boundary conditions. By applying the third boundary con¬ dition, we have 00 6fiURt(2 + t cos 6) . a V , /A • a « ,, o,s e»(2 + ,»)(l + «cosg)»sin 9 - 2, A- W 8in me = 0 (4-21) m — 1 To make the two terms equal to each other, it is necessary to expand the first term, p(0), into a Fourier series. By recalling the useful relation f+ 2aTos90 + a2 = 5} sin m6 1 we shall transform p(0) into a corresponding form, rewriting it as /n\ — SpUR \ e , e ~\ . a VW ej(2 + t!) [i + e cos 9 + (1 + e cos 0)2J Sm We thus have with - = a 6 Z a 00 tsin9 20 sin 9 = o V ia- sin me 1 + t cos 6 1 + 2a cos 6 + a2 Lj 1 By differentiating the last equation with respect to €, 00 / c sin 0 \ _ sin 0 _ V' ( — l)m~!am sin md de \1 + c cos e) ~ (1 + e cos 0)2 ~~ M e \/l — e2 1 By combining the last two expressions, we obtain the Fourier series for p(0) as * oo P(9) = C»(2t+^) X (-1)““‘a" [l + (l -S)»\ sin m6 (4'22) 1 where a = €. i + Vi-s
78 Theory of Hydrodynamic Lubrication By substituting Eq. (4-22) into Eq. (4-21) and equating coefficients in sin md, we obtain (L\ _ 12iiUR(— 1)—'a" [ m ] , 23) " (2) _ C2(2 + t2) [ + (1 + «2)WJ ( ' This expression, together with A2(z) = — KiAi(z), enables us to evaluate ^'(2) “ c' COsh KlD ~ C'-(2 + «’) f1 + (1 - »“)»] A.fy - -K,C, co.li K,i - - [l + ji-^] from which by dividing one into the other K\ — ^ ~ (4-24) Kl ~ [1 + (1 - e2)*]2 { } The last boundary condition is satisfied by the form chosen for p(z). Thus Eq. (4-17) with Am(z) given by the recurrence formula (4-20) constitutes a solution of Eq. (4-15) with the boundary conditions as given by Eq. (4-16). It can be shown that the series of Eq. (4-17) converges for the entire bearing domain and for 0 < e < 1.0. The load components are given by W sin <f> = jy*/2 f*r p(0,z) sin 6 R dd dz TT.cos <f> = — jy*/2 JqW p(0>z) cos BRdddz By using Eq. (4-17) in the above expression and noting the orthogonality of sin md and cos ra0, we have W cos 0 = 0 or $ and W-W.[ ^ S = (2 + «»)(! - «»)» 127r2e{ 1 - [tanh (K2L/D))/[K2L/D\\ K } where = \2yU RhLnr/C2{2 + €2)(1 — e2)^ is the load capacity of an infinitely long bearing and K22 = 2(1 -f- Kie)/{2 — Kit). As L/D—> <*>, Eq. (4-25) becomes equal to the load capacity of the infinite bearing. The frictional drag equation is
Incompressible Lubrication; Finite Bearings 79 By integrating the first term directly and the second term by parts, we obtain „ 2rcpURL ^ Ct CL'2 f2' . C(1 - e2)W ± 2R J-L/2 Jo p( sm eRdedz The integral above is the expression for load capacity evaluated previously, and thus in terms of the coefficient of friction the drag is given as — / = *2 h i (4-27) C J 6e{ 1 — [tanh (KJj/D))/[KJj/D]) “2 ^ '' The oil flow out of the z — L/2 end of the bearing is by Eq. (1-136) f2* h3 dp Q‘ jo 12m dz R dd L 2 which integrates to a + H-) (4.M) This analysis duplicates the results obtained for the infinitely long bearing in that the negative pressures are a mirror image of the positive pressures and the locus of the shaft center is a straight line at right angles to the applied load. It is of interest to note that Eq. (4-28) is the same as Eq. (2-36), i.e., it consists only of the flow contributed by the supply pressure pB. The pressure gradients being the same and opposite over 0 to x and x to 2x, there is no contribution of flow owing to the finiteness of the bearing. Numerical Methods The nonanalytical methods employed for the solution of the Reynolds equation range over the entire field of numerical analyses. They include ordinary mathematical relaxation, electrolytic tank models, d-c ana¬ logues, and digital computers. The following paragraphs deal with some of the techniques employed; this is followed by a tabulation of the avail¬ able results. 4-3. Digital Computers. By using the substitutions x z X=D Z=L c _ h - _ V (C\ 2 C P nN\Rj we have for the dimensionless Reynolds equation
80 Theory of Hydrodynamic Lubrication 12 3/ 7+1,/ hi /-i 4;/-, />/.! ?•/ A/ A/ 4-., Fig. 4-2. Grid element for computer solutions. Referring to Fig. 4-2, we have in finite increments for the three terms of the above equation 1%3 P*. 3+1 V' i _ jU3 Ax iJ- Pi.i — Pi, 3-1 * Ax Ax ^3 Pi+l.j Pi.j ^3 P*'./ Pi-l.i Az •-H.i A2 Az d/l /it, /+V£ j—Yi dx Ax from which we have by solving for pitj n /&». i-H ~ I (D\ r L3 Pi+l.) I L3 ‘ \L/ L <+M,i Az2 + AZ2 J Ax + P.-./ = ft 3 P*> 3+1 i ^ 3 PiJ-l “i.i+H Ax2 ^ Ax2 /D\2 M+Vi,i + hi j \L/ Az2 '+‘ i+H "b M Ax2 (4-30) which is of the form Pi,i = «o + aip»+i.; + CL2Pi-\,j + «3 P«.y+i + a*Pi,i-1
Incompressible Lubrication; Finite Bearings 81 with a0, ai, a2, a3, and a4 given constants for each point (i,j) of the mesh and the pressure p,., a function of these constants and the four surrounding pressure. For n X m points in the mesh there will be n X m simul¬ taneous equations which can be solved either in matrix form or by an iteration process. If the latter is used, the iterative process is repeated until an error smaller than the prescribed value A is reached. The error is defined by i-it-i < A 11 y-it-i 9/16 3/4 3/4 3/4 3/4 3/4 3/4 3/4 3/4 9/16 9/16 where k is the number of iterations performed. The allowable error A has to be kept to very small values, in the order of small fractions of 1 per cent, to obtain good accuracy. The individual pressures can, of course, differ from iteration to iteration by an amount larger than A. The pressure distribution is not much affected by the density of the grid, but it affects the numerical integration for load capacity. There are two ways of improving the accu¬ racy of the load capacity with a given number of points: one is to fit in additional pressure points and then perform the ordinary step integration; the other is to perform trapezoidal integration. The latter method can be shown to be achieved if the ordi¬ nary integration is used but by assign¬ ing to each boundary point a multi¬ plying factor of % and to each corner point a multiplying factor of %6 as shown in Fig. 4-3. points remain at their original value. Of course, Eq. (4-30) can also be used for ordinary mathematical relaxation by using the various techniques described in works on numer¬ ical analysis. 4-4. Electrical Analogues. The basic fact upon which the electrolytic tank and d-c analogue methods are based is that the differential equation governing the relation between electric potential and current is the same as for pressure and fluid flow. In general the flow of current in the x and z directions can be written as 3/4 3/4 3/4 3/4 3/4 3/4 3/4 3/4 9/16 Fig. 4-3. Multiplying factor for trape¬ zoidal integration. All internal . _ HdE u k dx HdE k dz
82 Theory of Hydrodynamic Lubrication where k is the resistivity and H is the area of the conductor. Referring to Fig. 4-4, which represents an element of electrolytic substance, with the lower surfaces nonconducting, we have q 11 4- h — 13 — I a + Iv = ix dz + iz dx + — (ix dz) dx — ix dz — iz dx ox Q — — (iz dx) dz + Iy = 0 dz By using the expression from above, A (K + A (K = = L dx\k dx) dz\kdz) y A which is the same differential equation as that of Reynolds with (4-31) 6 V dh dx In Eq. (4-31) k is the resistivity, H is the height of the electrolyte, and I is the current per element of area A. Thus, if an electrolytic bath in which the above relations are fulfilled is set up, the potential at any point in the electrolyte will correspond to the pressure in the bearing. With k and y. constant the above relations read H « h* Ik a it dh A~&liUTx A model as shown in Fig. 4-5, where the depth of the electrolyte is lyte. Fig. 4-5. Electrolytic tank analogue. proportional to hz and where current is fed at each element of area A proportional to 6yU dh/dx, will generate potentials at the nodal points proportional to the bearing pressures. If the bearing has a line of sym¬ metry, only half of the bearing surface is used, the center line being made
Incompressible Lubrication; Finite Bearings 83 a nonconductor. The boundaries are kept at potentials corresponding to the boundary pressures, mostly zero. The 24-in. long model described in Ref. 5 had the bottom made of wood coated with celluloid. There was a tendency for the electrolyte to split at hmin, and the surface had to be roughened there; this tendency limited the solutions to e < 0.8 when the depth of the bath was only 0.032 in. The bath consisted of a weak solution of potassium dichromate in dis¬ tilled H20 with a resistivity of 870 ohms/in. at 80°F. Side plates and electrodes were chromium-plated copper, and the current was taken from 110 volts 60 cycles a-c source reduced to 6 volts by a transformer. This avoided the difficulties encountered with direct current, which set up stray potentials and currents. The depth of immersion of the electrodes is unimportant, but the resistivity of the elec¬ trolyte has to be uniform throughout the bath and should be of a low value. A d-c network is similar to the electrolytic tank in that pressures are simulated by volt¬ ages and flow by current but differs in that the resistance to flow is represented not by the depth of an electrolyte but by resistors. Con¬ sidering now a current Iv flowing into the junction point of four resistors (Fig. 4-6), we have from I — E/R E0 Fig. 4-6. Element of resist¬ ance network. Iv = 11 + It. + Iz + Ia — or E 0 = Ei Eq — Ei E0 R1 R2 1 Rz Iy -f- E1/R1 -b E2/R2 + Ez/Rz -b Ea/Ra Ez Eq - Ea Ra (4-32) 1/Rl + l//?2 + l/#3 “H 1/R\ Equation (4-32) expresses the voltage E0 of any point in the field as a function of its neighboring potentials and the constants IV) Ri, R2, Rz, and Ra and is of the same form as Eq. (4-30) with _ fhj'j+M Ax Iv = 6ir Ri _ /L V Az2 ~ \d) h^j etc. In all the methods described above, the immediate results are a pressure distribution for a given 0, e, </>, and L/D ratio. In order to obtain values for load capacity, flow, and friction, numerical integration has to be used. These expressions are Ex = — ^ Y Pi.j cos 6itj Ax Az i-it-i m n Ft = ^ ^ pi,j sin $i,j Ax Az (4-33 o) (4-336) j = 1 f -1
84 Theory of Hydrodynamic Lubrication where for a fixed coordinate system (x,y) the line 0 = 0 must coincide with the y axis. The resultant force is, of course, given by The flow is given by Eq. (1-136). If we relate this to a flow coefficient q expressed by The last term of Eq. (4-36) is multiplied by L over the entire span of the bearing. This, as explained in the preceding chapter, is incorrect, since after 02 striation takes place and the bearing is not covered with lubricant over the entire length L. From 02 on, the bearing is covered with lubri¬ cant over an accumulated length L' < L, which can be readily calculated. Since from 02 on the axial pressure gradients are zero, no more side leakage occurs, and thus the flow inside the bearing is constant from 02 to 03. From continuity where V is the accumulated extent of the lubricant filmlets and h2 is the F = y/Fx2 + Fv2 and the attitude angle 4> is tan <f> — - tt 1/ (4-34) and using the substitutions of Eq. (4-29) we get for q (4-35) da where ~ is the slope of the pressure curves at the sides of the bearing. For frictional losses we have by Eq. (1-18) ^ IJh film thickness at the point where the film breaks down. From the above Thus Eq. (4-36) should be written
Incompressible Lvbrication; Finite Bearings 85 In the following section, values of S, e, qiu, q„ and / will be tabu¬ lated for a number of bearing designs. These results when used in the expressions will provide values of load capacity, minimum film thickness, flow, and friction as functions of bearing operating conditions and geometry. Tables 4-1 to 4-6 present the basic solutions of journal bearings obtained mostly on digital computers. These include full (ungrooved), partial, axial groove, and noncircular bearings, the latter being made up of several nonconcentric circular lobes. No separate results are given for bearings with circumferential grooving, for these grooves essentially convert such a bearing into two or more bearings of reduced L/D ratios, and solutions to these can be obtained from the results given. The boundary conditions upon which the solutions are based are essentially those described under assumption 3 at the opening of the chapter, i.e., as given by Eq. (3-15). Any other particular design, such as grooved bearings with eccentric loading or multiple-lobe bearings with unequal lobes, can be evaluated by the use of the fundamental solutions presented. A vast number of significant curves and cross plots can be drawn from the assembled material; only a few of these will be presented in the following brief discussions. 4-5. Full Journal Bearings. The data for full bearings are given in Table 4-1, and, according to the boundary conditions of Eq. (3-15), con¬ tain p = 0 at 0 = 0. This is not always the case with full journal bear¬ ings; for the lubricant can be admitted at any angle 0i. In fact, even if the admission point were located at 0 = 0 for some given operating condition, it would not remain at 0 = 0 for any other operating condition, since the 0 = 0 line shifts with each change of <f>. However, in practice this 0i is usually not very far from the 0 = 0 line, and a full bearing is not very sensitive to shifts even of the order of ±20°. This can be verified from the data of Table 4-2, where the performance of 180° bearing is seen to differ little from a full 360° bearing. If 0i is moved deep into = C( 1 - €) Qi. = q^NDLC Q. = q, | NDLC F = fW (4-38) JOURNAL BEARING SOLUTIONS
86 Theory of Hydrodynamic Lubrication Table 4-1. Full Journal Bearings L/D € S Qb CR/Of Po/P 0o - a* 0, - a* 00 0 00 0 1.0 71 148 0.1 0.247 0 3 69 139 0.2 0.123 0 2.57 1.22 7 67 128 0.3 0.0823 0 1.90 1.24 11 64 118 0.4 0.0628 0 1.53 1.31 15 62 108 0.5 0.0483 0 1.32 1.36 18 58 98 0.6 0.0389 0 1.20 1.51 21 54 87 0.7 0.0297 0 1.10 1.63 22 49 76 0.8 0.0211 0 0.962 2.01 22 42 62 0.9 0.00114 0 0.721 2.76 18 32 45 0.95 0.000605 0 0.568 3.921 14 23 32 1 0 00 0.00 1.0 86 120 0.1 1.35 0.16 5 79 112 0.2 0.632 0.316 12.9 1.89 9 74 105 0.3 0.382 8.04 1.93 13 68 97 0.4 0.261 0.607 5.80 2.08 16 62 89 0.5 0.179 4.31 2.23 18 56 82 0.6 0.120 0.938 3.21 2.41 19 50 74 0.7 0.0765 2.36 2.69 19 43 64 0.8 0.0448 1.24 1.71 3.15 18 36 53 0.9 0.0191 1.38 1.06 4.10 12 25 39 0.95 0.00855 1.49 0.675 5.32 7 16 27 X 0 oo 0.00 88 108 0.1 4.30 0.187 6 81 100 0.2 2.01 0.376 40.9 1.97 11 75 93 0.3 1.235 25.7 2.11 14 68 86 0.4 0.785 0.750 17.11 2.25 16 62 79 0.5 0.497 11.95 2.41 17 55 71 0.6 0.320 1.12 8.08 2.76 17 48 64 0.7 0.185 5.48 3.13 16 41 56 0.8 0.0920 1.50 3.25 3.72 15 33 47 0.9 0.0312 1.69 1.59 4.67 10 23 35 0.95 0.0119 1.79 0.869 6.27 6 15 24 Vs 0 00 1.00 90 102 0.1 9.36 7 83 95 0.2 4.42 87 2.02 12 75 87 0.3 2.72 56 2.21 15 68 80 0.4 1.67 35.5 2.36 16 62 73 0.5 1.04 23.8 2.50 16 54 66 0.6 0.642 15.9 2.94 16 47 59 0.7 0.354 10.1 3.38 15 40 51 0.8 0.166 5.6 4.02 13 32 43 0.9 0.0498 2.37 5.18 9 22 32 0.95 0.0170 1.18 7.04 6 15 22
Incompressible Lubrication; Finite Bearings Table 4-1. Full Journal Bearings (Continued) 87 L/D e S 9* (R/C)f Po/P do — a* 02 - a* X 0 00 0.00 1.00 90 98 0.1 15.9 0.196 8 83 92 0.2 7.58 0.393 153 2.03 13 75 84 0.3 4.69 98.5 2.26 15 68 77 0.4 2.85 0.782 61.4 2.40 16 61 70 0.5 1.78 40.0 2.64 16 54 63 0.6 1.07 1.10 26.7 2.97 16 47 56 0.7 0.591 16.6 3.51 15 39 48 0.8 0.266 1.56 8.93 4.20 13 31 40 0.9 0.0738 1.77 3.49 5.56 8 21 29 0.95 0.0231 1.86 1.58 7.30 6 15 20 X 0 00 1.00 90 95 0.1 25.3 9 83 89 0.2 11.7 234 2.05 14 75 82 0.3 7.07 150 2.27 15 68 75 0.4 4.42 94 2.50 16 61 68 0.5 2.77 61 2.80 16 54 61 0.6 1.63 40.3 3.04 16 47 53 0.7 0.887 28.7 3.56 15 39 46 0.8 0.389 13.1 4.27 12 31 38 0.9 0.104 5.0 5.75 8 21 28 0.95 0.0304 2.06 7.53 6 15 20 X 0 1.00 90 94 0.1 35.5 9 83 88 0.2 16.8 337 2.05 14 75 81 0.3 10.3 214 2.24 15 68 73 0.4 6.73 135 2.46 16 61 66 0.5 3.80 88 2.67 16 54 60 0.6 2.31 57.3 3.06 16 47 52 0.7 1.225 34.6 3.50 15 39 45 0.8 0.544 18.1 4.38 12 31 36 0.9 0.1415 6.57 5.93 8 21 27 0.95 0.0395 2.62 7.88 6 15 20 X 0 1.00 90 94 0.1 47.0 9 83 87 0.2 22.5 450 2.03 14 75 80 0.3 13.9 285 2.28 15 68 73 0.4 8.49 180 2.47 16 61 65 0.5 5.23 119 2.72 16 54 60 0.6 3.10 77 3.05 16 47 52 0.7 1.64 46 3.50 15 39 44 0.8 0.725 24 4.40 12 31 36 0.9 0.184 8.6 6.02 8 21 26 0.95 0.0506 3.3 8.18 6 15 20 * In full journal bearings the load angle a is arbitrarily defined as the angle between the line of centers 00' and the load line, measured in the direction of rotation. See Figs. 3-1 and 3-6.
88 Theory of Hydrodynamic Lubrication Table 4-1. Full Journal Bearings (Continued) L/D € S CR/C)f Po/P 00—01 4> 02 — Ot H 0 1.00 90 94 0.1 63.8 9 83 87 0.2 29.6 599 2.04 14 75 80 0.3 17.9 371 2.24 15 68 73 0.4 11.0 236 2.44 16 61 66 0.5 6.85 155 2.74 16 54 59 0.6 4.19 101 3.15 16 47 52 0.7 2.17 60 3.64 15 39 44 0.8 0.935 30.9 4.49 12 31 36 0.9 0.235 10.8 6.16 8 21 26 0.95 0.0626 4.08 8.35 6 15 20 Table 4-2. Centrally Loaded Partial Journal Bearing L/D Item ft c' 0.1 0.2 0.4 0.6 0.8 0.9 0.97 0.1 0.2 0.4 0.6 0.8 0.9 0.97 0.1 0.2 0.4 0.6 0.8 0.9 0.97 0.1 0.2 0.4 0.6 0.8 0.9 0.97 Values for <f>, S, vin. Q*, or (ft/C)/when arc Bpan 0, deg, is 360 0.240 0.121 0.0628 0.0410 0.0224 0.0111 0.96 0.90 0.72 0.50 0.24 0.13 4.8 2.57 1.52 1.20 0.961 0.756 180 73 61 50 43 33 26 15 0.347 0.181 0.0884 0.0530 0.0261 0.0132 0.00384 0.965 0.89 0.70 0.485 0.24 0.12 0.038 3.55 2.01 1.29 1.06 0.855 0.681 0.416 120 0.877 0.432 0.181 0.0841 0.0340 0.0147 0.00406 0.96 0.87 0.68 0.47 0.24 0.12 0.037 6.02 3.26 1.78 1.21 0.853 0.653 0.399 60 49 32 23 17 15 11 5.75 2.66 0.931 0.322 0.0755 0.0241 0.00495 0.95 0.87 0.66 0.43 0.23 0.12 0.037 19.7 10.1 4.67 2.40 1.10 0.667 0.372
Incompressible Lubrication; Finite Bearings 89 Table 4-2. Centrally Loaded Partial Journal Bearing (Continued) l/d - n* Item « Values for 4>, 8, g,a, <z*. or (R/C)f when arc span 0, deg, is 360 150 100 75 45 30 * 0.2 62 58 57 50 50 0.4 51 40 33 33 0.6 42 32 31 20 20 0.8 29 25 22 15 15 0.9 23 20 17 5 0.2 0.414 0.435 1.33 1.93 7.174 22.0 0.4 0.175 0.177 0.454 0.65 2.38 7.75 0.6 0.0875 0.0943 0.160 0.213 0.8212 2.26 0.8 0.0318 0.0370 0.0512 0.064 0.158 0.369 0.9 0.0159 0.0167 0.00914 0.003 7in 0.2 0.88 0.875 0.875 0.874 0.4 0.935 0.78 0.67 0.662 0.6 0.78 0.57 0.49 0.445 0.438 0.8 0.54 0.37 0.30 0.24 0.227 0.9 0.415 0.24 0.217 g* 0.2 0.165 0.10 0.05 0.0196 0.0086 0.4 0.290 0.135 0.08 0.027 0.012 0.6 0.350 0.145 0.09 0.0213 0.011 0.8 0.32 0.140 0.075 0.0222 0.0105 0.9 0.29 0.115 0.067 L/D - 1 Item 0.1 0.2 0.4 0.6 0.8 0.9 0.97 0.1 0.2 0.4 0.6 0.8 0.9 0.97 Values for <p, 8, gin, g*, or (R/C)f when arc span 0, deg, is 360 180 150 120 100 75 60 45 30 79 74 63 50 36 26 15 1.33 0.631 0.264 0.121 0.446 0.0188 0.00474 78 69 59 45 32 24 15 1.40 0.670 0.278 0.128 0.0463 0.0193 0.00483 0.714 0.275 0.125 0.0410 0.0190 72 58 44 36 27 22 13 2.14 1.01 0.385 0.102 0.0531 0.0208 0.00498 1.33 0.206 0.059 0.023 2.33 0.82 0.28 0.097 0.025 68 51 34 25 18 15 11 8.52 3.92 1.34 0.450 0.101 0.0309 0.00584 7.97 2.64 0.882 0.167 23.5 7.52 2.35 0.383 0.081
90 Theory of Hydrodynamic Lubrication Table 4-2. Centrally Loaded Partial Journal Bearing (Continued) L/D * l Item Values for 4>, S, $in, it, or (R/Of when arc span fi, deg, is 360 180 150 120 100 75 60 45 30 0.1 1.07 1.10 1.05 1.01 0.2 1.13 1.13 1.02 0.98 1.00 0.92 0.888 0.874 0.4 1.26 1.15 1.00 0.90 0.75 0.725 0.677 0.667 0.6 1.37 1.07 0.865 0.73 0.65 0.86 0.51 0.457 0.443 0.8 1.46 0.86 0.675 0.51 0.42 0.36 0.29 0.252 0.235 0.9 1.60 0.70 0.538 0.36 0.32 0.24 0.17 0.125 0.97 1.58 0.52 0.23 0.074 it 0.1 0.16 0.147 0.0892 0.0261 0.0288 0.0132 0.2 0.319 0.277 0.27 0.156 0.135 0.080 0.0431 0.0402 0.0185 0.4 0.607 0.472 0.43 0.238 0.18 0.14 0.0598 0.0380 0.0180 0.0 0.938 0.590 0.56 0.273 0.21 0.155 0.0630 0.0352 0.0174 0.8 1.24 0.605 0.46 0.267 0.215 0.13 0.0571 0.0135 0.9 1.38 0.556 0.43 0.232 0.14 0.093 0.0480 0.97 1.49 0.466 0.178 0.0339 R -f 0.1 26.4 14.1 14.5 29.1 cJ 0.2 12.8 7.15 7.44 1 14.8 0.4 5.79 3.61 3.60 :::::::::::: 6.61 0.6 3.22 2.28 2.10 i 3.29 0.8 1.70 1.39 1.27 1.42 0.9 1.05 0.921 0.855 0.822 0.97 0.514 0.483 0.461 0.422 L/D - M 0.1 0.2 82 75 80 75 69 72 64 63 61 56 53 49 50 0.4 61 58 53 48 46 40 37 33 33 0.6 48 45 42 38 32 31 27 25 25 0.8 33 31 24 28 26 22 20 17 17 0.9 0.97 0.1 0.2 24 14 4.31 2.03 23 14 4.38 2.06 22 21 13 5.42 2.51 15 17 1 16 11 14.2 13 2.10 3.12 4.29 6.47 11.62 29.3 0.4 0.779 0.794 0.80 0.914 1.05 1.50 2.14 3.80 9.94 0.0 0.319 0.321 0.31 0.354 0.395 0.51 0.695 1.16 2.73 0.8 0.0923 0.0921 0.082 0.0973 0.102 0.135 0.149 0.224 0.464 0.9 0.97 0.1 0.2 0.0313 0.00609 1.12 1.21 0.0314 0.00625 1.11 1.19 0.031 0.0324 0.00631 1.05 1.06 0.033 0.036 0.0422 0.00704 1.00 0.090 1.00 0.99 0.922 0.725 0.83 0.4 1.40 1.28 1.11 1.00 0.95 0.84 0.755 0.698 0.677 0.6 1.58 1.28 1.03 0.892 0.76 0.65 0.552 0.498 0.470 0.8 1.77 1.16 0.976 0.695 0.61 0.43 0.334 0.279 0.255 0.9 0.97 1.81 1.92 1.03 0.855 0.814 0.542 0.380 0.40 0.33 0.211 0.104 0.137
Incompressible Lubrication; Finite Bearings 91 Table 4-2. Centrally Loaded Partial Journal Bearing (Continued) Item 0.1 0.2 0.4 0.6 0.8 0.0 0.97 0.1 0.2 0.4 0.6 0.8 0.9 0.97 360 0.194 0.384 0.775 1.15 1.55 1.70 1.80 85.6 40.9 17.0 8.10 3.26 1.60 0.610 L/D - « Values for S, gin. ?«. or (R/C)f when aro span fi, deg, is 180 0.185 0.360 0.648 0.852 0.934 0.905 0.812 44.0 21.6 9.96 5.41 2.54 1.38 0.587 150 0.32 0.56 0.715 0.725 0.695 120 0.130 0.238 0.386 0.472 0.475 0.427 0.312 36.6 18.1 8.20 4.43 2.17 1.24 0.550 100 75 0.21 0.33 0.34 0.38 0.27 0.135 0.20 0.23 0.205 0.18 60 0.0488 0.0815 0.121 0.130 0.117 0.098 0.0676 48.6 24.2 10.3 4.93 2.02 1.08 0.490 0.0527 0.0743 0.0838 0.0708 30 0.0256 0.0357 0.0412 0.0362 0.0290 L/D - K 0.1 82 81 77 72 0.2 75 74 70 66 65 60 59 52 52 0.4 61 59 57 51 49 42 41 34 34 0.6 47 45 43 40 38 34 30 28 28 0.8 31 30 29 28 26 23 22 19 19 0.9 22 21 20 21 20 18 17 0.97 12 13 12 11 s 0.1 16.2 16.3 18.4 35.8 0.2 7.57 7.60 7.94 8.45 9.52 12.4 16.0 24.61 48.7 0.4 2.83 2.84 2.86 3.04 3.33 4.0 5.20 8.03 15.6 0.6 1.07 1.08 1.06 1.12 1.33 1.50 1.65 2.36 4.44 0.8 0.261 0.263 0.256 0.268 0.27 0.31 0.333 0.935 0.74 0.9 0.0736 0.0736 0.074 0.0743 0.080 0.077 0.0844 0.97 0.0101 0.0104 0.0105 0.0110 _ 0.1 1.10 1.10 1.06 1.01 0.2 1.20 1.19 1.19 1.08 1.06 0.965 0.925 0.896 0.4 1.39 1.31 1.26 1.19 1.04 0.97 0.817 0.732 0.730 0.6 1.59 1.35 1.27 1.02 0.94 0.77 0.630 0.552 0.509 0.8 1.78 1.30 1.14 0.85 0.72 0.56 0.413 0.327 0.280 0.9 1.88 1.19 1.00 0.702 0.55 0.43 0.285 .0.97 1.95 1.05 0.538 0.102 dm 0.1 0.182 0.193 0.154 0.0674 VI 0.2 0.396 0.380 0.36 0.309 0.26 0.18 0.175 0.0853 0.046 0.4 0.787 0.700 0.67 0.525 0.43 0.28 0.192 0.120 0.065 0.6 1.19 0.942 0.86 0.612 0.53 0.34 0.217 0.149 0.0806 1 0.8 1.57 1.19 0.93 0.641 0.50 0.33 0.205 0.131 0.0715 G.9 1.77 1.07 0.81 0.593 0.425 0.28 0.176 0.97 1.97 1.01 0.502 0.127 R — / 0.1 322 163 124 121 ■cf 0.2 153 79.4 60.4 58.7 0.4 61.1 35.1 26.6 24.5 • 0.6 26.7 17.6 13.5 11.2 0.8 8.80 6.88 5.65 4.27 0.9 3.50 2.99 2.63 2.01 0.97 0.922 0.877 0.832 0.713
92 Theory of Hydrodynamic Lubrication the converging region, then the case can be treated as being equivalent to a partial bearing and appropriate results can be obtained from Table 4-2 with 0 = $2 — di being the equivalent arc. The value of friction, however, will be higher than for the equivalent partial bearing; for it must include the shearing losses over the zero-pressure region of the full bearing. M id) Fig. 4-7. (a) Loci of shaft center and points of maximum pressure in full journal bearings; (6) locus of points where pressure wave ends in full journal bearings. Num¬ bers refer to L/D ratio. Figure 4-7 shows the loci of angles where the maximum pressure, the minimum film thickness, and the end of the pressure wave occur. It is seen that angles <f> and 0O depend very little on the L/D ratio; for low eccentricities, they are practically identical, and for high eccentricities their divergence is less than 10°. However, the value of 02 is appreciably reduced by a drop in the L/D ratio. 4-6. Centrally Loaded Partial Bearings. Table 4-2 gives the perform¬ ance of partial bearings where the load vector passes through the mid¬ point of the bearing arc, or a/(3 = A sample plot of the Sommerfeld
Incompressible Lubrication; Finite Bearings 93 <00 V E 2 1.0 -o B E ” 0.10 0.01, Bearing arc 0, deg , v/0 10 Fig. 4-8. Plot of Sommerfeld number vs. Fig. 4-9. Locus of shaft center for partial arc span. Solid line, e — 0.2; dashed journal bearings, a/t3 =* L/D = 1; line, € =* 0.8; numbers refer to L/D ratio, numbers refer to bearing arc. number versus bearing arc in Fig. 4-8 illustrates the previously mentioned fact that the load capacity is not appreciably affected by reducing the bearing arc from 360° to 180° or even 150°. The variation of the shaft locus with arc span 0 is shown in Fig. 4-9. 4-7. Eccentrically Loaded Partial Bearings. The results tabulated in Table 4-3 are the basic data for any bearing analysis. They give the result¬ ing hydrodynamic forces as a function of any arbitrary combination of param¬ eters 0, L/ Df €, and </>. The solutions presented in preceding and following sections were prepared from this basic information by a proper summation of the force vectors resulting for the various bearing elements. As they stand, the data constitute solutions for partial bearings in which the load is at any arbitrary position with respect to the bearing arc, or with a/0 as a vari¬ able. A sample variation of the locus of shaft center with a change of a is shown in Fig. 4-10. 4-8. Axial-groove Bearings. These bearings are made up of a series of concentric partial bearings separated by axial grooves for admitting Fig. 4-10. Variation of attitude with load angle. 0 = 120°; L/D = 1; numbers refer to a/0 ratio.
94 Theory of Hydrodynamic Lubrication Table 4-3. Eccentrically Loaded Partial Journal Bearings 0 = 150° L/D ( a/0 <t> S 9in l X 0.2 0.200 20 4.16 0.020 0.266 35 1.80 0.038 0.380 48 0.782 0.105 0.493 61 0.477 0.186 0.546 83 0.510 1.09 0.206 0.612 103 0.738 1.14 0.169 0.859 156 32.7 1.18 0.0098 0.3 0.273 34 0.91 0.061 0.4 0.286 32 0.520 0.735 0.085 0.453 47 0.244 0.882 0.234 0.600 65 0.231 1.14 0.407 0.646 98 0.389 1.26 0.347 0.866 155 24.6 0.5 0.100 10 6.58 0.0049 0.213 23 0.614 0.046 0.293 31 0.308 0.103 0.353 37 0.224 0.105 0.420 42 0.164 0.248 0.567 50 0.133 0.417 0.6 0.300 30 0.176 0.553 0.118 0.443 38 0.114 0.708 0.266 0.660 66 0.125 1.17 0.607 0.683 93 0.259 1.35 0.534 0.860 156 23.5 1.56 0.030 0.7 0.100 10 1.22 0.006 0.220 22 0.160 0.059 0.320 28 0.101 0.131 0.397 30 0.0793 0.212 0.467 35 0.0690 0.308 0.520 37 0.0630 0.386 0.8 0.213 23 0.0695 0.287 0.061 0.334 25 0.0497 0.142 0.725 86 0.183 1.40 0.722 0.800 156 24.1 1.76 0.040 0.9 0.253 17 0.0225 0.0628 0.368 20 0.0187 0.146 0.453 22 0.0172 0.236 0.540 24 0.0161 0.361 0.95 0.396 15 0.0077 0.147 0.493 16 0.0071 0.247
Incompressible Lubrication; Finite Bearings 95 Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) 0 = 150° L/D e «/0 * S Qin 1 <7* 1 0.2 0.100 10 71.2 0.002 0.200 25 4.92 0.031 0.273 34 2.15 0.069 0.380 48 1.05 0.148 0.487 62 0.756 0.241 0.552 72 0.780 0.408 0.274 0.620 102 1.08 0.367 0.230 0.860 156 36.8 0.0659 0.0145 0.4 0.207 24 1.32 0.0565 0.360 41 0.435 0.671 0.215 0.500 52 0.286 1.00 0.385 0.610 74 0.327 0.810 0.545 0.655 97 0.568 0.732 0.467 0.866 155 27.2 0.115 0.029 0.5 0.100 10 7.02 0.007 0.207 24 0.732 0.081 0.293 31 0.368 0.155 0.360 36 0.277 0.231 0.427 41 0.228 0.342 0.500 46 0.209 0.94 0.425 0.6 0.214 23 0.383 0.497 0.079 0.394 36 0.163 0.733 0.303 0.500 41 0.137 0.865 0.450 0.672 64 0.176 1.19 0.810 0.694 91 0.371 1.10 0.720 0.807 155 26.3 0.130 0.044 0.7 0.100 10 1.31 0.009 0.220 22 0.181 0.088 0.317 27 0.116 0.197 0.340 29 0.111 0.232 0.390 31 0.0972 0.307 0.407 35 0.0840 0.445 0.500 36 0.835 0.776 0.461 0.550 38 0.0788 0.600 0.8 0.233 20 0.0797 0.316 0.090 0.333 25 0.0581 0.428 0.202 0.523 32 0.0468 1.04 0.500 0.754 52 0.0785 1.55 1.07 0.740 84 0.255 1.17 0.97 0.866 155 27.3 0.285 0.059 0.9 0.360 21 0.0213 0.218 0.423 21H 0.01905 0.308 0.486 22 0.01875 0.410 0.814 43 0.0414 1.63 1.16
96 Theory of Hydrodynamic Lubrication Table 4-3. Eccentrically Loaded Partial Journal Bearings 0Continued) 0 -= 150° LID € a/0 0 S qm q* 1 0.95 0.394 16 0.00825 0.222 0.420 17 0.00910 0.264 0.474 19 0.00980 0.364 0.534 20 0.0105 0.478 0.2 0.180 28 9.80 0.049 0.240 39 4.77 0.103 0.346 53 2.85 0.202 0.574 59 2.05 0.339 0.567 80 2.17 1.19 0.359 0.643 99 2.89 1.20 0.307 0.827 161 54.0 1.19 0.0252 0.3 0.246 38 2.63 0.120 0.4 0.160 21 2.68 0.67 0.060 0.266 35 1.34 0.812 0.202 0.480 53 0.807 0.55 0.526 56 0.794 0.60 0.632 70 0.905 1.365 0.707 0.686 72 1.46 1.42 0.614 0.866 155 41.0 1.38 0.054 0.5 0.0932 11 8.77 0.016 0.196 25 1.26 0.120 0.266 35 0.798 0.241 0.333 40 0.642 0.365 0.406 44 0.559 0.505 0.560 51 0.521 0.752 0.6 0.500 41 0.303 1.03 0.622 0.706 59 0.420 1.21 1.05 0.734 85 0.897 1.09 0.422 0.876 154 39.7 1.18 0.081 0.7 0.0932 11 1.685 0.021 0.213 23 0.319 0.158 0.300 30 0.232 0.376 0.376 33 0.210 0.485 0.460 36 0.189 0.693 0.580 38 0.1815 0.940 0.8 0.180 18 0.159 0.331 0.105 0.327 26 0.105 0.360 0.500 30 0.0912 0.976 0.752 0.786 47 0.156 1.56 1.37 0.776 79 0.567 1.75 1.25 0.866 155 41.0 0.256 0.109
Incompressible Lubrication; Finite Bearings 97 Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) jS = 150° L/D ( a//3 S Qxa 9* 0.9 0.253 17 0.0226 0.063 0.306 20 0.0321 0.386 * 0.460 21 0.0307 0.588 0.540 24 0.0161 0.362 0.814 43 0.0417 1.15 0.95 0.334 15 0.01135 0.402 0.500 15 0.01065 0.618 x 0.2 0.156 31 25.1 0.068 0.220 42 15.3 0.125 0.326 56 9.62 0.243 0.453 67 8.00 0.335 ' 0.500 70 7.75 0.360 0.580 78 7.70 1.21 0.398 0.670 95 9.80 1.22 0.346 0.870 155 115 1.19 0.042 0.4 0.173 29 6.40 0.136 0.240 39 4.18 0.256 0.370 49 3.09 0.482 0.480 63 2.84 0.680 0.653 67 3.03 1.42 0.795 0.720 87 4.07 1.44 0.687 0.873 154 87.0 1.40 0.086 0.6 0.187 27 1.74 0.201 0.273 34 1.265 0.388 0.426 41 1.075 0.730 0.487 47 1.065 0.882 0.735 55 1.27 1.61 1.16 0.767 80 2.70 1.67 1.03 0.8 0.0866 12 0.954 0.04 0.220 22 0.319 0.257 0.326 26 0.277 0.496 0.506 29 0.257 0.947 0.830 42 0.398 1.66 1.40 0.820 73 1.51 1.85 1.34 0.875 154 86.0 1.80 0.173 0.9 0.100 10 0.154 0.042 0.246 18 0.0836 0.271 0.370 19.5 0.0788 0.515 0.526 21 0.0720 0.983 0.870 35 0.159 1.90 1.74
98 Theory of Hydrodynamic Lubrication Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) 0 = 120° e a/0 <f> S Qin g* CR/C)S 0.1 0 0 00 0.90 0 oo 0.190 17 36.4 0.910 0.00514 268 0.360 37 4.70 0.945 0.0350 33.5 0.438 48 2.85 0.97 0.0595 20.1 0.482 72 2.25 1.00 0.0809 15.5 0.510 79 2.12 1.03 0.0930 14.2 0.532 96 2.27 1.06 0.0940 14.8 0.554 114 2.74 1.07 0.0830 17.4 0.568 122 3.18 1.08 0.0735 19.9 0.586 130 3.85 1.08 0.0618 23.9 0.615 136 4.98 1.08 0.0487 30.6 0.651 142 6.88 1.07 0.0368 42.1 0.20 0 0 00 0.800 0 oo 0.190 17 13.8 0.818 0.0100 108 0.367 36 1.89 0.888 0.0685 14.8 0.451 47 1.20 0.936 0.116 9.19 0.502 60 0.998 1.00 0.160 7.32 0.532 76 1.00 1.05 0.184 6.98 0.552 94 1.15 1.11 0.189 7.54 0.570 111 1.48 1.15 0.170 9.18 0.581 120 1.77 1.16 0.150 10.7 0.597 130 2.20 1.17 0.126 12.9 0.622 135 2.90 1.17 0.100 16.8 0.656 142 4.09 1.15 0.0750 23.4 0.4 0 0 00 0.600 0 oo 0.192 18 3.33 0.635 0.0192 32.3 0.383 34 0.574 0.762 0.129 5.68 0.484 41 0.401 0.855 0.219 3.79 0.547 54 0.369 0.971 0.310 3.32 0.579 70 0.426 1.09 0.368 3.33 0.594 88 0.569 1.20 0.382 3.90 0.602 108 0.844 1.29 0.345 5.10 0.607 117 1.07 1.32 0.310 6.09 0.616 127 1.39 1.34 0.263 7.58 0.634 135 1.91 1.34 0.208 10.0 0.664 140 2.80 1.30 0.154 14.2 0.6 0 0 oo 0.402 0 00 0.196 17 0.837 0.45 0.0282 11.0 0.408 31 0.200 0.615 0.178 2.77 0.526 37 0.155 0.747 0.306 2.03 0.605 47 0.160 0.911 0.445 1.84 0.634 64 0.219 1.10 0.547 2.08 0.637 84 0.355 1.28 0.580 2.68
Incompressible Lubrication; Finite Bearings 99 Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) 0 = 120° t a/0 * S Qin (R/C)f 0.6 0.631 105 0.620 1.42 0.530 3.76 0.630 115 0.834 1.47 0.478 4.62 0.633 124 1.14 1.50 0.407 5.86 0.646 132 1.63 1.51 0.322 7.86 0.670 139 2.46 1.46 0.238 11.3 0.7 0 0 oo 0.300 0 00 0.199 14 0.376 0.352 0.0311 6.10 0.425 29 0.112 0.531 0.178 1.96 0.553 34 0.0925 0.682 0.342 1.52 0.643 47 0.0998 0.870 0.505 1.41 0.667 79 0.155 1.08 0.630 1.69 0.660 81 0.287 1.30 0.678 2.32 0.646 103 0.550 1.48 0.627 3.39 0.641 113 0.766 1.54 0.564 4.21 0.641 123 1.08 1.57 0.477 5.40 0.651 132 1.57 1.59 0.379 7.28 0.673 139 2.40 1.54 0.294 10.6 0.8 0 0 00 0.200 0 oo 0.204 16 0.142 0.255 0.0333 3.11 0.450 26 0.0563 0.443 0.212 1.37 0.587 29 0.0494 0.610 0.376 1.12 0.691 37 0.0559 0.820 0.668 1.05 0.706 56 0.104 1.17 0.790 1.38 0.685 78 0.230 1.33 0.785 2.04 0.660 101 0.495 1.54 0.725 3.11 0.651 112 0.714 1.61 0.652 3.91 0.648 123 1.03 1.67 0.560 5.06 0.655 132 1.53 1.68 0.442 6.88 0.676 140 2.38 1.62 0.326 10.0 0.9 0 0 OO 0.100 0 oo 0.216 15 0.0360 0.151 0.0336 0.130 0.490 21 0.0210 0.346 0.224 0.861 0.640 24 0.0197 0.529 0.404 0.756 0.762 39 0.0232 0.766 0.630 0.697 0.755 50 0.0626 1.05 0.815 1.09 0.712 75 0.182 1.35 0.892 1.81 0.675 99 0.449 1.60 0.800 2.89 0.662 110 0.673 1.69 0.890 3.69 0.656 122 1.00 1.73 0.945 4.81 0.660 131 1.51 1.76 0.975 6.58 0.679 138 2.38 1.70 0.905 9.67 0.97 0 0 00 0.0300 0 00 0.243 11 0.00594 0.0706 0.0311 0.500
100 Theory of Hydrodynamic Lubrication Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) 0 - 120° L/D € a/0 4> S 9>n (R/C)f 1 0.97 0.552 14 0.00493 0.270 0.227 0.455 0.712 15 0.00487 0.462 0.417 0.428 0.854 19 0.00588 0.715 0.665 0.374 0.803 44 0.0372 1.02 0.872 0.878 0.733 73 0.152 1.36 0.965 1.66 0.685 98 0.420 1.62 0.890 2.77 0.669 110 0.649 1.73 0.807 3.56 0.660 121 0.982 1.79 0.687 4.68 0.663 130 1.50 1.82 0.547 6.42 0.681 138 2.39 1.75 0.403 9.47 'A* 0.10 0 0 00 0.900 0 00 0.177 19 60.0 0.915 0.0098 427 0.324 41 10.7 0.96 0.0534 76.7 0.400 52 7.22 1.00 0.0864 50.7 i 0.467 63 5.78 1.03 0.116 39.5 1 0.510 79 5.46 1.06 0.131 36.5 i 0.544 94 5.86 1.08 0.131 38.1 0.578 111 7.10 1.16 0.122 44.9 0.629 125 9.74 1.10 0.0877 60.3 0.691 137 15.9 1.08 0.0557 97.3 0.20 0 0 00 0.802 0 00 0.178 19 22.0 0.825 0.0195 171 0.333 40 4.19 0.916 0.107 33.5 0.416 51 3.02 0.98 0.169 22.9 0.492 61 2.52 1.05 0.231 18.4 0.537 76 2.56 1.11 0.264 17.6 0.569 92 2.95 1.15 0.262 19.2 0.599 108 3.82 1.18 0.233 23.4 0.641 123 5.55 1.19 0.176 32.5 0.698 136 9.48 1.15 0.111 54.2 0.4 0 0 00 0.600 0 00 0.180 18 i 5.22 0.650 0.0380 50.5 0.355 38 | 1.26 0.825 0.209 12.3 0.456 46 0.977 0.95 0.342 9.05 0.548 54 1 0.902 1.08 0.450 7.63 0.595 69 1.04 1.21 0.522 7.89 0.619 86 1.41 1.31 0.532 9.38 i i 0.636 104 2.11 1.36 0.470 12.5 0.664 120 3.47 1.38 0.358 18.6 0.709 135 6.50 1.31 0.228 32.9 * The data for L/D = % and L/D = Y\ of 0 = 120° are taken from as yet unpub¬ lished results by J. Boyd and A. A. Raimondi, Westinghouse Research Laboratories.
Incompressible Lubrication; Finite Bearings 101 Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) 0 - 120° L/D € a/0 4> S 9io 9* (R/C)f X 0.6 0 0 00 0.402 0 00 0.185 18 1.28 0.468 0.0544 16.8 0.389 31 0.414 0.723 0.300 5.56 0.509 39 0.354 0.909 0.776 4.38 0.618 46 0.359 1.11 0.663 3.85 0.661 63 0.499 1.30 0.775 4.38 0.671 80 0.834 1.46 0.800 5.87 0.672 100 1.50 1.55 0.716 8.71 0.684 118 2.79 1.57 0.550 14.0 0.718 134 5.73 1.46 0.348 26.1 0.70 0 0 00 0.300 0 00 0.190 17 0.562 0.375 0.061 9.07 0.412 31 0.2207 0.344 0.177 3.69 0.544 36 0.197 0.88 0.555 3.02 0.660 41 0.210 1.12 0.757 2.68 0.697 56 0.332 1.35 0.922 3.27 0.697 77 0.646 1.53 1.04 4.78 0.689 97 1.31 1.63 0.838 7.57 0.693 116 2.61 1.66 0.646 12.6 0.721 134 5.58 1.53 0.406 24.3 0.8 0 0 00 0.199 0 00 0.197 17 0.202 0.281 0.062 4.42 0.444 27 0.101 0.602 0.280 2.32 0.587 29 0.0942 0.85 0.628 1.96 0.712 35 0.105 1.13 0.658 1.75 0.739 52 0.205 1.36 1.02 2.38 0.724 74 0.494 1.60 1.07 3.91 0.705 95 1.14 1.72 0.961 6.68 0.701 116 2.47 1.75 0.742 11.7 0.725 134 5.50 1.61 0.467 23.0 0.9 0 0 00 0.100 0 00 0.212 15 0.0482 0.18 0.0675 1.72 0.493 21 0.0323 0.53 0.41 1.24 0.648 22 0.0315 0.81 0.688 1.09 0.784 25 0.0371 1.13 0.973 0.965 0.789 45 0.11 1.43 1.17 1.63 0.754 70 0.369 1.66 1.20 3.20 0.722 94 1.00 1.81 1.085 5.97 0.709 115 2.38 1.84 0.837 10.9 0.729 133 5.53 1.69 0.528 22.1 X 0.1 0 0 00 0.90 0 00 0.157 22 139 0.92 0.015 994 0.291 45 33.2 0.97 0.0675 237
102 Theory of Hydrodynamic Lubrication Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) 0 = 120° L/D e a/(3 <t> S Qin (R/C)f hi 0.1 0.368 53 23.9 1.01 0.100 169 0.451 66 19.5 1.03 0.133 134 0.511 79 18.5 1.07 0.152 124 0.556 94 19.8 1.09 0.152 128 0.603 108 23.9 1.10 0.132 150 0.665 120 32.0 1.10 0.102 197 0.727 133 48.8 1.08 0.0682 298 0.2 0 0 00 0.805 0 00 0.158 21 50.6 0.835 0.029 395 0.301 44 13.3 0.94 0.135 104 0.388 53 10.1 1.01 0.201 76.6 0.480 62 8.61 1.18 0.290 62.3 0.543 74 8.61 1.14 0.305 59.2 0.567 92 9.85 1.18 0.303 63.7 0.627 105 12.6 1.20 0.268 77.5 0.680 118 18.0 1.19 0.203 105 0.734 133 29.0 1.12 0.132 165 0.40 0 0 OO 0.600 0 00 0.162 20 11.9 0.665 0.0595 115 0.329 41 3.9 0.885 0.271 37.9 0.434 47 3.24 1.01 0.402 29.6 0.548 58 3.00 1.15 0.530 25.0 0.612 67 3.42 1.27 0.605 25.3 0.646 82 4.55 1.35 0.605 29.7 0.670 100 6.83 1.39 0.532 39.8 0.706 115 10.9 1.37 0.412 58.3 0.747 130 19.6 1.31 0.277 99.0 0.60 0 0 OO 0.40 0 00 0.169 20 2.87 0.494 0.087 37.5 0.370 36 1.24 0.81 0.398 16.4 0.498 40 1.14 1.02 0.602 13.7 0.631 45 1.12 1.21 0.782 11.6 0.689 57 1.51 1.40 0.900 12.7 0.707 75 2.53 1.52 0.905 17.2 0.713 95 4.64 1.59 0.800 26.3 0.729 113 8.60 1.56 0.622 42.5 0.757 129 17.0 1.46 0.418 77.5 0.7 0 0 oo 0.300 0 oo 0.175 19 1.28 0.408 0.100 19.9 0.399 32 0.638 0.775 0.462 10.4 0.539 35 0.611 1.01 0.688 8.92 0.678 38 1 0.618 1 1.26 0.922 7.44
Incompressible Lubrication; Finite Bearings 103 Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) 0 - 120° L/D t a/0 <t> S Qia {R/C)f H 0.7 0.729 53 0.950 1.46 1.05 8.78 0.735 73 1.87 1.61 1.05 13.2 0.731 93 3.94 1.69 0.935 22.1 0.739 113 8.0 1.65 0.719 38.1 0.762 129 16.6 1.54 0.486 71.6 0.8 0 0 oo 0.200 0 00 0.185 18 0.433 0.322 0.112 9.25 0.438 28 0.277 0.74 0.522 6.02 0.588 29 0.266 1.01 0.800 5.15 0.734 33 0.282 1.29 1.04 4.29 0.773 48 0.54 1.51 1.19 5.72 0.764 69 1.36 1.70 1.20 10.1 0.748 90 3.36 1.79 1.06 18.8 0.748 110 7.40 1.76 0.835 34.2 0.9 0 0 00 0.10 0 00 0.205 16 0.0935 0.228 0.12 3.23 0.496 20 0.0750 0.70 0.578 2.64 0.657 22 0.0745 1.01 0.880 2.34 0.806 24 0.0825 1.32 1.17 1.88 0.823 42 0.253 1.59 1.35 3.33 0.794 65 0.955 1.79 1.37 7.57 0.766 88 2.87 1.88 1.21 16.0 0.756 110 6.98 1.85 0.95 31.3 0.770 127 16.1 1.69 0.63 63.9 0 = 100° L/D < a/0 <t> S q\n 9* 1 0.2 0.11 9 142 0.0016 0.28 22 6.85 0.803 0.0222 0.43 23 1.95 0.889 0.0752 0.51 64 1.32 0.997 0.142 0.52 68 1.32 0.218 0.148 0.54 86 1.42 1.08 0.159 0.56 114 2.05 1.16 0.141 0.65 135 5.00 1.18 0.0716 0.70 156 37.6 1.18 0.0151 0.79 171 42.5 1.19 0.0107 0.92 168 83.5 1.20 0.0012 0.3 0.11 9 69.0 0.709 0.00235 0.28 22 3.29 0.743 0.0328
104 Theory of Hydrodynamic Lubrication Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) 0 - 100° L/D e a/0 S 0in g« 1 0.3 0.44 36 1.02 0.841 0.111 0.53 57 0.745 0.966 0.201 0.56 84 0.925 1.09 0.238 0.57 113 1.49 1.23 0.213 0.65 135 3.94 1.27 0.111 0.95 165 645 1.29 0.0019 0.4 0.53 47 0.476 0.278 0.55 55 0.485 0.263 0.6 0.12 8 7.1 0.424 0.00475 0.29 21 0.472 0.351 0.52 33 0.191 0.657 0.234 0.54 36 0.184 0.695 0.265 0.56 39 0.180 0.738 0.295 0.62 78 0.40 1.40 0.487 0.92 113 3.79 1.38 0.193 0.93 107 614 1.59 0.0038 0.8 0.12 8 0.827 0.231 0.0059 0.31 19 0.097 0.300 0.0727 0.42 23 0.0688 0.368 0.143 0.48 37 0.0595 0.430 0.205 0.54 26 0.0572 0.470 0.245 0.67 63 0.245 1.22 0.652 0.63 107 0.782 1.61 0.605 0.64 122 1.40 1.71 0.48 0.66 129 2.20 1.72 0.363 0.95 165 710 1.77 0.0051 0.9 0.13 7 0.133 0.136 0.00604 0.36 16 0.0279 0.203 0.0734 0.50 20 0.0219 0.304 0.179 0.55 20 0.0211 0.342 0.212 0.70 70 0.187 1.22 0.736 0.64 106 0.725 1.66 0.685 0.65 125 1.72 1.80 0.475 0.66 139 2.19 1.81 0.412 0.95 165 710 1.88 0.0057 0.95 0.36 14 0.0101 0.150 0.0685 0.45 15 0.00945 0.198 0.116 0.50 15 0.00898 0.224 0.142 0.52 15 0.00895 0.235 0.153 0.65 105 0.704 1.71 0.728 0.64 121 1.33 1.72 0.470 0.65 122 1.47 1.79 0.542
Incompressible Lubrication; Finite Bearings 105 Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) 0 = 100° L/D M 0.2 0.3 0.4 0.5 0.6 «/0 0.13 0.16 0.40 0.52 0.53 0.55 0.58 0.66 0.79 0.85 0.91 0.11 0.27 0.41 0.48 0.57 0.59 0.67 0.78 0.92 0.11 0.26 0.34 0.43 0.52 0.54 0.55 0.59 0.61 0.63 0.67 0.74 0.77 0.78 0.98 0.27 0.58 0.62 0.77 0.11 0.20 0.28 0.50 0 7 24 40 68 72 85 112 134 141 130 169 9 23 39 62 83 111 133 152 192 9 24 31 37 48 51 55 81 109 122 133 146 148 152 198 23 52 108 153 9 15 22 35 183 12.1 4.31 2.98 3.02 3.27 4.76 10.00 56.5 127 945 87.7 5.75 2.22 1.70 2.10 3.40 7.76 32.3 835 40.5 2.92 1.75 1.25 1.04 1.04 1.06 1.50 2.73 4.04 6.85 16.4 20.9 29.2 740 1.57 0.677 2.34 28.7 8.62 1.75 0.795 0.392 gin 0.808 0.809 0.926 1.07 1.08 1.14 1.20 1.20 1.20 1.19 1.08 0.710 0.769 0.89 1.02 1.20 1.30 1.18 1.30 1.29 0.617 0.692 0.701 0.851 0.981 1.06 1.27 1.60 1.42 1.40 1.40 1.40 0.59 1.00 1.42 1.47 0.429 0.44 0.53 0.805 g« 0.0033 0.00395 0.121 0.224 0.232 0.246 0.214 0.119 0.0225 0.0137 0.0027 0.0048 0.059 0.18 0.312 0.368 0.324 0.181 0.057 0.0041 0.0064 0.0775 0.146 0.236 0.356 0.352 0.413 0.492 0.435 0.350 0.273 0.120 0.096 0.075 0.0054 0.095 0.503 0.535 0.090 0.0094 0.045 0.111 0.383
106 Theory of Hydrodynamic Lubrication Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) 0 = 100° L/D e a/0 S q\n 9* H 0.6 0.53 37 0.382 0.843 0.428 0.56 39 0.376 0.906 0.484 0.61 49 0.414 1.00 0.596 0.64 76 0.855 1.38 0.725 0.65 106 2.05 1.59 0.656 0.71 139 9.17 1.605 0.265 0.73 142 11.75 1.60 0.221 0.78 152 27.7 1.56 0.113 0.7 0.29 21 0.37 0.42 0.125 0.665 114 1.70 1.68 0.87 0.8 0.22 8 1.0 0.250 0.0121 0.305 20 0.146 0.364 0.136 0.41 24 0.113 0.489 0.261 0.49 26 0.103 0.529 0.301 0.53 27 0.100 0.656 0.428 0.70 40 0.14 0.98 0.78 0.70 71 0.488 1.51 0.99 0.66 104 1.69 1.785 0.87 0.67 118 3.021 1.845 0.715 0.68 127 4.593 1.76 0.555 0.69 131 5.724 0.496 0.79 151 29.6 1.75 0.157 0.9 0.12 8 0.159 0.132 0.0126 0.22 13 0.0561 0.163 0.055 0.33 17 0.0394 0.271 0.14 0.50 15 0.0330 0.317 0.55 20 0.0327 0.505 0.375 0.76 54 0.0575 0.985 0.844 0.70 70 0.392 1.56 1.12 0.70 85 0.785 1.61 1.08 0.68 102 1.55 1.90 1.00 0.78 152 27.7 1.84 0.171 0.92 192 745 1.89 0.0123 0.95 0.36 14 0.0131 0.214 0.131 0.45 15 0.0125 0.287 0.204 0.50 15 0.0122 0.331 0.248 0.52 15 0.0121 0.349 0.266 0.71 84 0.721 1.80 1.145 0.70 100 2.14 1.915 1.03 0.67 113 2.32 1.99 0.926 0.68 117 2.91 1.985 0.835 0.68 119 3.17 1.98 0.80 0.79 151 29.8 1.89 0.18
Incompressible Lubrication; Finite Bearings 107 Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) 0 = 100° LID t a/0 <t> S <?« n 9* H 0.2 0.05 105 1180 1.19 0.005 0.10 10 294 0.811 0.006 0.23 27 30.1 0.511 0.055 0.36 44 12.9 0.953 0.148 0.51 71 9.54 1.10 0.271 0.52 73 9.60 1.11 0.278 0.56 84 10.1 1.16 0.286 0.61 109 14.4 1.21 0.250 0.69 131 25.2 1.31 0.147 0.83 157 161 1.21 0.032 0.3 0.09 11 138 0.719 0.009 0.23 27 14.4 0.715 0.052 0.38 42 6.7 0.732 0.222 0.59 81 6.31 0.853 0.428 0.63 107 10.3 0.977 0.375 0.81 119 14.5 1.25 0.322 0.91 169 944 1.30 0.0078 0.4 0.20 10 65.8 0.622 0.0118 0.24 26 7.50 0.723 0.108 0.31 34 4.85 0.809 0.194 0.395 40.5 3.78 0.905 0.290 0.53 52 3.28 1.085 0.47 0.55 55 3.25 1.115 0.49 0.63 77 4.47 1.31 0.57 0.65 105 8.27 1.42 0.50 0.67 118 12.1 1.425 0.405 0.71 129 19.1 1.42 0.296 0.78 147 52.0 1.36 0.137 0.79 151 65.4 1.41 0.110 0.91 169 942 1.40 0.0105 0.6 0.20 10 13.9 0.438 0.0177 0.35 25 1.75 0.58 0.16 0.52 38 1.17 0.955 0.595 0.56 39 1.12 0.995 0.59 0.68 72 2.42 1.46 0.86 0.68 102 6.00 1.025 0.75 0.74 136 25.1 1.66 0.38 0.75 140 30.1 0.29 0.91 169 945 1.60 0.016 0.8 0.11 21 1.57 0.0255 0.29 21 0.349 0.931 0.205 0.40 25 0.285 0.596 0.37 0.48 27 0.276 0.726 0.50
108 Theory of Hydrodynamic Lubrication Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) 0 - 100° L/D € a/0 <t> S j qin I i H 0.8 0.52 28 0.270 0.690 0.464 0.74 66 1.27 1.50 1.13 0.71 99 4.74 1.525 1.00 0.72 114 8.61 1.856 0.816 0.73 123 13.2 1.857 0.677 0.91 191 945 1.80 0.021 0.9 0.11 9 0.24 0.155 0.0238 0.33 17 0.081 0.352 0.221 0.50 20 0.0933 0.589 0.458 0.55 30 0.0754 0.680 0.550 0.78 62 0.862 1.66 1.27 0.73 97 0.428 1.94 1.13 0.95 195 1,010 1.89 0.023 0.95 0.37 13 0.0234 0.285 0.203 0.46 14 0.0224 0.407 0.326 0.49 16 0.0229 0.452 0.371 0.73 99 4.08 1.99 1.19 0.73 107 6.50 2.015 1.05 0.72 113 8.32 2.035 j | 0.985 0 = 75° IK 0.2 0.167 10 91.0 0.00155 0.00155 0.207 12 41.2 0.0290 0.0290 0.514 59 1.87 0.0933 0.0548 0.54 73 1.87 0.147 0.0655 0.56 78 1.95 0.110 0.0665 0.687 151 20.6 0.0628 0.01485 0.715 154 27.1 0.0907 0.01175 0.3 0.136 8 101 0.0111 0.0111 0.179 10 41.0 0.023 0.023 0.515 49 15.8 0.0853 0.0755 0.535 67.5 1.18 0.203 0.0915 0.540 72 1.23 0.205 0.0945 0.660 148 11.6 0.111 0.0294 0.686 151 15.8 0.096 0.0230 0.5 0.143 7.5 23.1 0.00182 0.00182 0.515 34 6.67 0.6365 0.0885 0.527 38 0.312 0.174 0.103 0.553 46 0.407 0.219 0.120 0.567 65 0.481 0.320 0.101 0.580 125 2.15 0.342 0.124 0.590 135 2.81 0.326 0.118 0.615 152 6.67 0.215 0.0012
Incompressible Lubrication; Finite Bearings 109 Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) P - 75° L/D € a/P 4> S 3in 9* IK 0.7 0.540 27 0.125 0.129 0.099 0.580 26 0.126 0.202 0.133 0.587 114 1.35 0.414 0.219 0.590 124 1.35 0.495 0.200 0.9 0.467 12 0.0254 0.480 17 0.0266 0.0842 0.0842 0.527 18 0.0245 0.1065 0.1065 0.620 101 1.04 0.757 0.430 1 0.2 0.167 10 62.5 0.00244 0.00244 0.207 12 46.5 0.00425 0.00425 0.512 59 2.29 0.1055 0.055 0.520 74 2.32 0.1625 0.094 0.527 78 2.40 0.163 0.0956 0.687 151 21.20 0.0685 0.0230 0.713 154 32.30 0.0557 0.0167 0.3 0.127 8 108 0.00175 0.00175 0.107 10 44.2 0.0036 0.0036 0.512 49 1.31 0.875 0.111 0.534 68 1.36 0.227 0.136 0.540 72 1.44 0.237 0.140 0.660 148 12.71 0.121 0.044 0.686 151 17.4 0.103 0.0346 0.5 0.512 34 0.487 0.670 0.129 0.560 56 0.550 0.310 0.201 0.567 60 0.578 0.337 0.215 0.621 141 6.03 0.270 0.1145 0.654 145 8.07 0.107 0.0927 0.7 0.528 28 0.158 0.160 0.145 0.554 34 0.147 0.305 0.194 0.567 113 1.14 0.562 0.308 0.570 123 2.40 0.529 0.275 0.9 0.474 17 0.0265 0.0842 0.0842 0.527 18 0.0245 0.1065 0.1065 0.620 101 1.03 0.720 0.430 0.635 155 0.84 0.719 0.442 K 0.2 0.200 12.5 60.8 0.0083 0.0083 0.495 54 4.28 0.162 0.1335 0.505 56 4.44 0.168 0.136 0.514 64 4.20 0.185 0.1495 0.526 78 4.48 0.207 0.161 0.586 139 12.30 0.145 0.0965
110 Theory of Hydrodynamic Lubrication Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) 0 = 75° L/D e «/0 <t> S 9>n 9* Vi 0.2 0.640 145 15.80 0.109 0.0666 0.727 153 44.20 0.064 0.0324 0.3 0.126 8 131 0.0034 0.0034 0.140 12 62.1 0.0067 0.0066 0.515 49 2.47 0.204 0.184 0.541 67 2.64 0.279 0.223 0.548 72 2.90 0.281 0.222 0.70 90 28.2 0.061 0.676 147 21.7 0.132 0.076 0.5 0.515 34 0.890 0.242 0.221 0.535 43 0.895 0.295 0.269 0.560 51 0.935 0.324 0.320 0.575 60 1.065 0.398 0.359 0.606 137 8.25 0.354 0.239 0.654 145 12.2 0.284 0.178 0.7 0.527 28 0.273 0.253 0.245 0.560 30 0.260 0.314 0.292 0.580 36 0.272 0.360 0.328 0.620 111 3.26 0.507 0.500 0.615 117 3.90 0.647 0.476 0.615 122 4.85 0.600 0.430 0.9 0.475 17 0.0365 0.161 0.161 0.581 19 0.0342 0.439 0.248 0.648 94 1.50 0.942 0.736 0.654 104 2.46 0.662 0.915 174 3,090 0.00414 0.00414 M 0.2 0.154 11 211 0.0082 0.180 14 123 0.0135 0.500 60 12.4 0.184 0.527 73 12.3 0.206 0.530 78 12.6 0.209 0.714 149 77.0 0.061 0.741 152 99.0 0.0487 0.3 0.126 8 202 0.0062 0.154 11 95 0.0123 0.500 50 6.95 0.243 0.540 67 7.15 0.298 0.547 72 7.50 0.307 0.694 146 50.2 0.113 0.714 149 63.4 0.092
Incompressible Lubrication; Finite Bearings 111 Table 4-3. Eccentrically Loaded Partial Journal Bearings (Continued) 0 = 75° L/D «/0 4 S ?in 9* H 0.5 0.515 34 2.46 0.300 0.573 55 2.69 0.455 0.580 59 2.90 0.203 0.660 138 2.82 0.225 0.7 0.515 34 0.680 0.573 36 0.683 0.646 110 0.885 • 0.650 120 12.5 0.9 0.460 18 0.785 0.251 0.520 19 0.762 0.305 0.672 92 4.06 0.93 0.666 98 5.10 0.905 0.880 171 1,085 0.0158 the lubricant, such as shown in Fig. 4-11. The solutions to these are obtained by simply adding vectorially the forces generated in each seg¬ ment. The performance of two-, three-, and four-groove bearings is given in Table 4-4. Fig. 4-11. Axial groove bearings, (a) Three-groove bearing; (6) four-groove bearing. 4-9. Noncircular Bearings. These bearings are composed of sections whose centers of curvature are not in the geometric center of the bearing. Thus although the individual lobes are circular, the geometric configura¬ tion of the bearing as a whole is not circular. Two such common designs are the elliptical, or the two-lobe, and the three-lobe bearings shown in Fig. 4-12. The distance between the bearing center 0 and the lobe centers, distance d, we have called ellipticity; the ratio d/C = 8 we have
112 Theory of Hydrodynamic Lubrication Fig. 4-12. Noncircular bearings, (a) Elliptical bearing; (b) three-lobe bearing. called the ellipticity ratio, where C is the difference in the radii between journal and bearing lobes. The analyses of these bearings, and thus all subsequent formulas, refer to and use this lobe clearance C as the operating variable. In these bearings, there are, even for a concentric shaft position, two reference clear¬ ances : a minor clearance given by an inscribed circle denoted by Cm and a major clearance given by a circle circumscribed on the bearing. There are also in these bearings two kinds of eccentricity ratios: one given by c = e/C or €m = e/Cm, called bearing eccentricity ratio, which is a measure of the distance of journal center from the geometric center of the bear¬ ing, and one given by €1,2,3 = eit2,3/C, called lobe eccentricity ratio, which is a measure of the distance of journal center from the centers of lobe 1, 2, or 3. Likewise, there are two kinds of attitude angles, a bearing and indi¬ vidual lobe attitude angles. These quantities can be expressed one in terms of the other, and we shall do that for each bearing separately. The trigonometry of the elliptical bearing is shown in Fig. 4-13. By denoting by 1 the lower lobe and by 2 the upper lobe, we have from simple trigonometry €1 = x^e2 S2 -f- 2(8 cos <t> c sin <f> €1 Fig. 4-13. Trigonometry of elliptical bearings. = sin- €2 = \/e2 + 82 — 2(8 cos <f> . . € sin <f> 4> 2 = sin-1 - €2 (4-39)
Incompressible Lubrication; Finite Bearings 113 Table 4-5 gives the performance of elliptical bearings for three elliptic- ity ratios, 5 = J4, 5 = H, and 6 = The case of 6 = 0 corresponds to a circular axial-groove bearing; 5 = 1 is the limiting case of a no-clear¬ ance bearing. In this table we have tabulated the larger of the two lobe eccentricity ratios, for this is the parameter that provides the value of the minimum film thickness = C( 1 - €ltl) (4-40) It is of interest to note that the journal under loading may actually travel upward, a result of the high initial convergence of the film. Thus, in Table 4-5 whenever the bearing attitude angle 0 is more than 90°, the lobe eccentricity ratio to be used in Eq. (4-40) is actually c2; for the mini¬ mum film thickness will occur in the upper lobe. Table 4-4. Axial Groove Bearings* L/D e Two-groove Three-groove Four-groove <t> 8 9* S * S IX 0.2 62 0.435 0.165 67 0.99 0.11 80 1.9 0.081 0.4 51 0.179 0.290 49 0.37 0.205 72 0.71 0.16 0.6 42 0.0943 0.350 36 0.15 0.29 49 0.245 0.29 0.8 29 0.037 0.320 27 0.05 0.39 25 0.259 0.38 0.9 23 0.0167 0.290 20 0.019 0.40 18 0.0222 0.37 1 0.2 64 0.714 0.27 69 1.33 0.16 79 2.04 0.12 0.4 53 0.275 0.43 50 0.48 0.30 72 0.87 0.6 45 0.125 0.56 37 0.187 0.425 49 0.30 0.39 0.8 28 0.041 0.46 26 0.059 0.51 25 0.07 0.51 0.9 22 0.019 0.43 20 0.022 0.66 19 0.025 0.53 X 0.2 64 2.10 0.32 72 3.00 0.245 80 4.23 0.20 0.4 53 0.80 0.56 54 1.06 0.47 72 1.7 0.42 0.6 42 0.31 0.715 39 0.385 0.68 49 0.55 0.67 0.8 24 0.082 0.725 27 0.103 0.92 25 0.11 0.84 0.9 22 0.031 0.695 20 0.033 1.04 19 0.035 0.90 X 0.2 70 7.94 0.36 74 9.61 0.30 79 11.6 0.272 0.4 57 2.86 0.67 56 3.30 0.625 72 4.5 0.56 0.6 43 1.06 0.86 41 1.17 0.90 49 1.3 0.89 0.8 27 0.256 0.93 28 0.27 1.17 25 0.23 1.15 0.9 20 0.074 0.87 20 0.0735 1.29 19 0.073 1.22 * For ungrooved bearing see Table 4-1. The trigonometry of the three-lobe bearing is shown in Fig. 4-14. The relations between bearing and lobe attitudes are, by referring to the figure,
114 Bottom lobe Right lobe Left lobe Theory of Hydrodynamic Lubrication €1 = \/e2 + 52 + 2e5 cos </> <f>i = sin' €l €2 -V (i + *) ^ « sill (t/3 + <#>) 02 sin O €2 (4-41) c3 = yjt <f> 3 = sin 2 + 52 — 2*5 cos — (fr'j l e sin (tt/3 — 0) Table 4-5. “Elliptical” Bearings* L/D Ui H 5 = K 5 = H Cm <t> Cl.2 S 9* <t> Cl,2 S Q* <t> Cl,2 S Qt 0.2 110 0.33 0.572 0.17 98 0.52 0.625 0.23 87 0.75 0.25 0.27 0.4 90 0.39 0.185 0.29 98 0.56 0.313 0.25 75 0.78 0.11 0.275 0.6 62 0.61 0.090 0.35 92 0.59 0.156 0.33 75 0.82 0.045 0.295 0.8 38 0.81 0.0333 0.305 70 0.74 0.051 0.33 75 0.835'0.034 0.30 0.9 32 0.90 0.0167 0.29 53 0.85 0.024 0.30 0.84 0.03 0.30 1.0 75 0.85 0.029 0.30 1.2 70 0.895 0.019 0.30 0.2 105 0.32 0.834 0.225 95 0.52 0.415 0.322 85 0.75 0.285 0.41 0.4 90 0.39 0.308 0.413 90 0.54 0.274 0.37 80 0.77 0.143 0.41 0.6 68 0.59 0.120 0.546 87 0.60 0.155 0.44 81 0.79 0.087 0.41 0.8 35 0.81 0.040 0.436 67 0.75 0.056 0.47 79 0.81 0.0607 0.44 0.9 30 0.90 0.019 0.430 48 0.86 0.024 0.43 78 0.84 0.0455,0.44 1.2 65 0.90 0.0202 0.45 0.2 100 0.31 2.0 0.275 90 0.51 1.3 0.52 80 0.76 0.53 0.68 0.4 90 0.39 0.57 0.53 85 0.54 0.67 0.57 80 0.78 0.26 0.69 0.6 62 0.61 0.29 0.72 83 0.61 0.32 0.67 75 0.82 0.085 0.72 0.8 38 0.81 0.071 0.72 65 0.76 0.11 0.72 75 0.83 0.067 0.73 0.9 30 0.90 0.0305 0.69 48 0.86 0.040 0.675 75 0.84 0.058 0.73 1.0 75 0.85 0.053 0.74 1.2 67 0.90 0.030 0.76 0.2 100 0.31 7.15 0.42 90 0.51 5.55 0.66 75 0.76 1.67 0.93 0.4 95 0.41 3.33 0.59 85 0.545 2.32 0.72 75 0.78 0.715 0.96 0.6 62 0.61 0.96 0.84 80 0.62 0.945 0.825 75 0.80 0.488 0.97 0.8 30 0.81 0.24 0.93 70 0.74 0.377 0.91 75 0.83 0.260 0.98 0.9 28 0.90 0.074 0.93 45 0.875 0.080 0.875 75 0.84 0.170 0.98 1.0 75 0.85 0.130 1.00 1.2 68 0.88 0.091 1.00 | 1 1 5 = % * For 5=0 see Table 4-4.
Incompressible Lubrication; Finite Bearings 115 Table 4-6. Three-lobe Bearings* L/D € 5 - lA 8 = H 4> s Qi <t> <i S Qi 1 0.2 42 0.58 0.45 0.125 50 0.71 0.21 0.18 0.4 53 0.63 0.18 0.185 50 0.76 0.12 0.185 0.6 55 0.71 0.10 0.20 50 0.81 0.071 0.20 0.8 50 0.815 0.048 0.235 45 0.86 0.039 0.21 1.0 30 0.965 0.0083 0.28 40 0.945 0.0095 0.23 M 0.2 45 0.57 0.84 0.265 50 0.71 0.43 0.355 0.4 55 0.63 0.40 0.31 52 0.76 0.21 0.35 0.6 55 0.71 0.20 0.335 50 0.81 0.11 0.38 0.8 50 0.815 0.084 0.425 45 0.86 0.054 0.41 1.0 30 0.965 0.011 0.50 40 0.945 0.0125 0.465 H 0.2 45 0.575 2.5 0.37 62 0.70 1.16 0.53 0.4 45 0.65 1.0 0.51 55 0.75 0.59 0.53 0.6 45 0.745 0.41 0.54 58 0.80 0.33 0.53 0.8 40 0.845 0.13 0.61 52 0.86 0.12 0.575 1.0 30 0.965 0.021 0.75 44 0.945 0.025 0.58 * For 5=0 see Table 4-4.
116 Theory of Hydrodynamic Lubrication Here too the basic clearance to be used in all formulas is C, the lobe clearance and the minimum film thickness is determined by the value of «i given in Table 4-6. The analyses for both of these bearings were conducted by fixing a bearing attitude angle 4> and eccentricity ratio c, calculating the various lobe attitudes by use of Eq. (4-39) or Eq. (4-41) and then obtaining the vertical and horizontal force components for each lobe from Table 4-3. A solution was obtained for 2Fx = 0, and then 2Fy provided the load capacity. The flow coefficient was obtained by summing the individual values for all lobes. FINITE THRUST BEARINGS Analytical Solutions As mentioned in Chap. 3, the analysis of thrust bearings is made easier by the simplicity of the expressions for film shape and by the simple boundary conditions p( 0) = p(£) = p(|) = p(-^) = 0 We may, as in the case of journal bearings, attempt to obtain an exact solution of the Reynolds equation for a simple slider. By writing h = ax for the film shape, we have from Eq. (1-12) S+IS+S-S-o <«2> The solution of Eq. (4-42) is assumed to be of the form = y«.(s)8in« Zv nx 1,3,5 the odd integers being due to symmetry in the z direction. By setting the length of the slider as it (the span being taken as a multiple of 7r), and since 1,3,5 7 for all 0 < z < 7T 4 — — we can rewrite Eq. (4-42) into
Incompressible Lubrication; Finite Bearings 117 By setting £ = nx and —!= k iraL 00 dp Bp V A dw» ™n\ • aI = na? = nZV{^_Fj8inn2 1 ^ sin rw: i d2p V' n2tt>n • a? “ ” 2/ T~sin 712 the differentiation of the series term by term being permissible because p(0) = p(r) = 0. By making all the coefficients of sin nz equal to zero, we have B2wn 1 Bwn ( 1 \ _k f d£2 £ d£ \ £2/ £2 (4-43) The homogeneous part of Eq. (4-43) is in the form of Bessel equations, and the nonhomogeneous part can be written as a series of £. Thus the complete solution of Eq. (4-43) is made up of wn = AnIi(£) + BnKi(£) 4- (C 4- Z)£ 4- E%2 4- * • *) and u>n = A'nIl(Q 4- B'uKt(t) 4- (C 4- D^~l 4- E^~2 + • • •) where h and K\ are the Bessel functions of the first and second kind. With the coefficients C, Z), E, . . . determined, the two series above become wn = AnIx(£) 4- BnKi(£) — k ^1 4- -|- 4- 7^7-32 + 7 . 52 32 + ' ' = A'nI,(£) 4- B'nKi(i) - k(f2 + 3 £"4 -b 5 • 32£"6 4- 7 • 52 32£~8 4_ . . .) Thus the solution of Eq. (4-43) can be written as follows: For small nx oo p(x,z) = |a./,(iw) + BnK^nx) 24/it/ [. . (nx)2 (nx)4 (nx)« 1) %+7W+ J) (4’44a) For large nx oo p(x,z) = J S-^ (^/.(nx) + B'nK\(nx) 24/.t/ f, , 3 , 5-32 , 7 • 52 32 , H xa2(na;)2 [ (nx)2 (nx)4 (nx)6 J! d2p 2 d^p = 2 y/l B2wn 2 dww 2u>, ax2 n a£2 71 Z/\£ a£2 £2 a£ £3
118 Theory of Hydrodynamic Lubrication The coefficients An and Bn must be determined from p(0) = p(x) = 0 for all values of z. This is done by making the terms in the bracket, which is independent of z, zero. However, their evaluation by analytical means is not easily obtained. The solution of these equations by numerical means is given later in the text. 4-10. The Step Bearing. A complete analytical solution of the step bearing can be obtained by essentially solving Laplace’s equation with one nonzero boundary condition. For the plane slider the film thickness is given by h = const, and Eq. (1-12) transforms into that of Laplace A solution of the following form is assumed: oo 1,3,5 1,3,5 with boundary conditions (see Fig. 3-16) p(0,z) = p(x, 0) = p(x,L) = 0 and 1,3,5 From the first two boundary conditions An = Cn = 0 From the last condition we have 1,3.5 1,3,5 from which sinh (mrBi/L) sinh (mrBi/L) 1,3,5 An identical analysis for region I yields
Incompressible Lubrication; Finite Bearings To evaluate pn, the requirement of continuity is imposed: 119 Uhi hi3_ dp ~2~ + 12pdx — -hi _i_ hih hz -Bi ” ~T + I2ii di Bt dp dx dp dx 2mr TVn 1,3,5 00 2 me Al mrB2 . mrz coth —7— sin L Jj Al_ mrBi . mrz coth —j— sin -j- When these expressions are used in the flow equations, a mrBi , , , mrB2\ . mrz 3 coth —■=— 4- hS coth 1 si 1,3,5 since -j— + h2z coth sin ^ = fyU(hi - h2) -y - hi) — sin ^ Z-/ mr L 1,3,5 00 2 1,3,5 4 . mrz — sin —=- = 1 mr L The above relation thus yields the value of pn: 24nUL(hi - h2) Pn = nV[hi3 coth (mrBi/L) + h2z coth (mrB2/L)] Thus Eq. (4-45) with the value of pn as given above provides the pres¬ sure distribution. The load capacity after proper integration is oo iv — 48mlJLz{hj — h2) V _1 tanh (mrBi/2L) + tanh (mrB2/2L) 7r4 Lj n4 h\z coth (mrBi/L) + h2z coth (mrB2/i 1.3.5 i/L) (4-46) The frictional force is F = nUL (Bi B2\ 2UUIJ(hi - h2y \ki V 7T3 00 X n’t*? 1 1.3.5 coth (mrBi/L) + h2z coth (mrB2/L)\ (4-47)
120 Theory of Hydrodynamic Lubrication Flow in at the leading edge is ^ UhiL 4:ULhi2(hi - h2) Qi = -t « n2 sinh (nrBi/L) [hi2 coth {mrBi/L) + A23 coth (mcB2/Li)] 1,3,5 Flow out at the trailing edge is „ UhJj . 4ULh22(hi - h2) Q* - — + ? 00 ^ n2 sinh {mcB2/L) [hi2 coth (mrBi/L) + /i23 coth (mrBi/L)] 1.3,5 The side leakage is then Qz = Qi - Q2 For a sectorial step bearing with d1 and d2 replacing B1 and B2 Laplace’s equation is a2p 1 d2p 1 dp _ dr2 r2 dd2 r dr By using r = Rxep r = Ri for p = 0 r = R2 for p = p2 = I11 ~ Hi the equation above becomes *2. 4. = n dp2 ^ dd2 The solutions to this equation are thus identical with those for a rec¬ tangular shape and are for region II oo / ax V Pn • ln (r/Ri) . , mrO = A -:-u sin T^/p-rp-x sinh i Li sinh [mrd2/ln {R2/Ri)\ ln (R2/Ri) In (rt2/fli) (4-50a) and for region I oo , flv _ V • 7l1r ln (r/^l) • U nr, } Z/ sinh [ftjr^/ln (ft,/Si) 1 Sm "In («2/fl,) S1" ln (ft2/fl,) (4-506) The expression for flow is given by . _ rco/i. h2 dp _ uRieph h2 dp 1 ~ T" + l2pr dd ” —2 + dd
Incompressible Lubrication; Finite Bearings 121 and by equating the two regions I and II, we have oo yK — pn (hi* coth + h23 coth sjn VUE = 6pwfti2e2p(/ti — hi) p2 \ p2 P2 / pi i By expansion of e2p into a Fourier series we obtain oo oo e2p — V Kn sin = V F — f e2p sin dpi sin 7— L^i pi LJ I pi Jo pi J pi n — 1 1 00 S2mc[(— l)n+1e2p* + 1] • nxp {nx)2 + (2p2)2 sin 7T i By solving for pn, we have \2pRJ<* ln (R2/Ri)[(-l)”+'(R2/Ri)2 + mi ~ hi) Pn { {nx)2 + [2 ln (Ri/Ri)}2\[hY coth [nxdi/ln {Ri/Ri)] + h2* coth [71x62/In (R1/R1)]} The total load capacity is n- 1 tanh + tanh nr9' ln {Rt/RxY ' v"“" ln WRQ3 } hi* coth j—- -N -f- h23 coth nT^2 ln (R2/R1) ' ' In (R2/R1) The torque required is (,„iy i \ (-1 )»+w+ft.j 1 L (mr)2 + [2 In («S//?,)]2J 1.3,5 1 hi3 COth y>y-r;-,-T + A.8 COth nT®2 In {Rt/RO ' ‘ ln (Rt/R,) (4-52) Flow becomes Qi = _ 2a> (ln A,3(A, - ^2)(«22 - tf,*) X {(wr)2 -I- [2 ln (^/fti)]2) sinh, _ . . n-1,3,5 In {Ri/ Rl) — (4-53) rnrdi hY coth p^-~T + h2* coth .—1 L ln (R2/R1) In (/?2//?i)J
122 Theory of Hydrodynamic Lubrication Q* = ^ + 2«ln(^-2) *,»(*, - *,)(*,* - «.2) 1 X ((n*)2 + [2 In (Rt/Rx)]2) sinh Y*? n-1.3.5 In {tl2/ Kl) [*‘*coth EM + w coth in (X/fio] (4‘54) 4-11. Slider with Exponential Film Shape. As is evidenced from pre¬ ceding paragraphs, the attainment of exact solutions to the Reynolds equa¬ tion is not easy even for simple configurations. However, as will be apparent from the following analysis, considerable simplification is introduced by expressing the film thickness by an exponential function. Its use can be justified by the fact that the performance of thrust bearings at constant viscosity is influenced very noticeably by the values of hi and h2 and much less by the actual film shape between the inlet and outlet ends. Mathematically, the simplification introduced by an expo¬ nential function is that of eliminating the variable coefficients in the two differential equations resulting from the homogeneous part of the assumed solution. By writing h = c&x Eq. (1-12) can be written as 3+S+»g-2^«-”' <«« The solution is assumed to be of the form p(x,z) = p(x) + q(x,z), and we have for the homogeneous equation 0 + *!+»£S-° with the boundary conditions now p(x) + q(x,z) = 0 on the edges of the bearing. By multiplying p(x) by e3bx and setting p(0) = p(B) = 0, we have for the particular solution , v 3mU \e~2bB - e-*>* , 1 - <r2W* _36x 2. "I P(I) = -PT [ 1 - + 1 - e~»‘ 6 " e \ (4~57) which is a solution to the infinitely wide slider with an exponential clearance. The homogeneous equation yields two differential equations Z" - \2Z = 0 X" + 3 bX' + A2X = 0 and the solutions to these two equations are of the form Z(z) = E cosh Az + F sinh Az X(x) = e~3bxl2(C cos nx + D sin nx) where n = \/\2 — (36/2)2 is a constant to be evaluated.
Incompressible Lubrication; Finite Bearings 123 Since we set p(0) = p(B) — 0, we have for boundary conditions of q(x,z) the following: Using q = 0 for x = 0 yields C = 0, and symmetry in z yields F = 0. This gives the homogeneous solution q(xyz) = 2(jne-36z/2 sin nx cosh \z From q(B) = 0, excluding the trivial case Gn = 0, we have The remaining boundary condition necessary to evaluate Gm is along z = L/2 for all x. By equating the two expressions, with z set equal to L/2, multiplying both sides by sin (kirx/B)} integrating from x — 0 to x = By and making use of the orthogonality properties, we have for Gm the following: r _ 6/z U (mir/B) m Bc2b\2 cosh XL/2 l(\ _ e-26B)[(_1)mg-36B/2 _ 1 ] -|_ (g-2 bB _ _ 1)me-36B/2 _ !] | 1 - or m = 1, 2, 3, . . . q(x,z) = -p(x) The final solution is thus ^ cosh \z + p(x) (4-58) m — 1 where p(x) is given by Eq. (4-57). The load capacity is, upon integration of Eq. (4-58),
124 Theory of Hydrodynamic Lubrication The friction is given by F = &k Ue-** - l) + 9(1 ~ e'tbB^] cb |_ 1 l) + 4(1 - e-•**) J 00 + c ^ Gm [«-»*/*(- 1)”‘ - 1] sinh ^ (4-60) m — 1 The flow into the bearing is n 4 - Se~bB - e-*bB T TT bB , c3e36*'2 V Gm(-l)mrrnr/B . , L\ Q“ - - 4(1 - UUe + “67" 2/ X Smh - 2 (4-61) and the side leakage 00 n c3 /wwA eZbB,2{ — l)m — 1 . ,XL Q‘ = 6^ 2/ W X Sinh ~2~ (4*62) The centroid is m* 1 X4 LV ' J)J 1 c263 I 4 [e-»*(36£ + 1) - 1](1 - e~2bB) b2B2(e~2bB - e~*bB) 9(1 - e~zbB) + 2(1 - e~zbB) (4-63) Figure 4-15 has some of the above results plotted for L/B = oo, 1, and These results differ from the solutions obtained for a slider with h = ax by less than 2 per cent. This reaffirms the previously expressed view that, once the ratio h\/h2 is fixed, the intermediate shape of the film is of no great significance. Numerical Solutions 4-12. Slider Bearing; Semianalytical Methods. The analytical ex¬ pression for pressure given by Eqs. (4-44) was arrived at by expanding both p and 6nU/aW in a Fourier series in z. This necessitates a new numerical evaluation of the power or Laurent series each time the value of nU/a2 is changed. This can be avoided by following the previous procedure of representing the solution as the sum of p(x), the solution for L/B = oo, and q(x,z), the solution to the homogeneous equation. Thus
Incompressible Lubrication; Finite Bearings 125 where the first right-hand term is clearly the particular solution identical with Eq. (3-50) and satisfies the boundary conditions p(x 1) = p(x2) = 0. t/2 >1 -CO e: l i i ! 1 r— / , 1 I h' / 1 / / V/ 7 0 20 40 60 80 pULtfW Fig. 4-15. Performance of exponential sliders. Numbers refer to L/B ratio. The homogeneous part is then written (4-M) n-1 where Un(x) = Ji(anx2)Yi(anx) - Yi(anx2)Ji(anx) Ji and Y\ being first-order Bessel functions of the first and second kind, respectively. Hyperbolic instead of trigonometric functions are used to improve the convergence of the results. The boundary conditions for Eq. (4-64) are q(xlfz) = q(x2,z) = 0 q(x'i) = q(x’~i) = p(x) which will be satisfied if Un(x) =0 at x = Xi and x = .r2 n tt //»\ x)(x2 x) » n( ) " ~(xi + X2)x By using these relationships to evaluate the constants and making use of the orthogonality properties of the involved functions, we have for the
126 Theory of Hydrodynamic Lubrication or "Y n&nJ U \ pressure distribution , , _ 6\iV \ (x, - x)(x - x2) t2 pM - -pr [ (*, + l2)x2 - 2 •0 7n^n«fl^(^n^l) Un(x) COSh OLnZ "j , . . L/ J12(«„X2) - Jl2(«nXi) COSh (anL/2) J n-1 n(s) / _ COSh ttng \1 x \ cosh (anL/2)/J (4-656) where = </i(a„x2) / T0(a„x) dx - Ti(a„x2) / '‘ Jo(anx) dx J XI J Xl The main difficulty arises in evaluating the condition t/„(xi) = J i(otnx2)Yi(anXi) — Y i(anx2)J i(anx2) = 0 As described in Ref. 16, the roots had to be evaluated to an extreme accuracy; for even the seventh decimal point could affect the results. 0.16 v(xt) =^fV n — 1 0 1.0 0 8 /hz V 8 Fig. 4-16. Load capacity of plane sliders. Numbers refer to L/B ratio. Also, the value of ynotnJ i(«„xi) had to be evaluated with a high degree of accuracy. The value of yn after performing the integrations yields - Jx{aai) [l -^//.(a„x2)]J where Hi is the Struve function of the first order.
Incompressible Lubrication; Finite Bearings 127 3.0 2.5 ■ 2.0 ■ € 1.5 ■ 1.0 - 0.5 - °0 Inflow — Side leokoge 0.25 0.5 /' 0.8 0.25 ,/6 % 0.5 2 0.8 1 e / s — 00 2 A — 6 CO 2 3 l/ht Fig. 4-17. Inflow and side leakage in plane Fig. 4-18. Coefficient of friction in plane sliders. Numbers refer to L/B ratio. sliders. Numbers refer to L/B ratio. The integration of Eqs. (4-65) yields 4 « + 1 (1 /«)(•> - 1) ^ ^ anynJi(anXi) j -W anx2\JtHt tanh anx2(a — l)(L/2B)\ (4-66) <xnx2[J i2(ocnx2) — J i2(a„Xi)] where the series term represents the correction due to the finiteness of the slider. The friction factor for the moving part is / _ 1 , pUB, a 2 + Wot* ® (4-67) The flow into the slider is UaLBa (, ir(a 1) l - a2 - 1 | (L/B)(a — 1) SotnynJi(otnXi)Ji(anx2) tanh gnx2(a - 1)(L/2£)| &nX*hJ T?(ot.nx2) J i2(a„:ri)] | (4-68) and out the sides icUaB2 (a - l)2 <xnynJi(anXi)[Ji(otnXi) — cul i(anx2)] tanh gnx2(a — \){L/2B) anX2[J i2(anx2) — J i2(an£i)) (4-69) Here, too, the final results, although based on analytical expressions, had to be given in numerical form. Figures 4-16 to 4-19 give the values of load capacity, friction, lubricant flow, and center of pressure with the
128 Theory of Hydrodynamic Lubrication constants of integration obtained with the use of a digital computer in a manner described in Ref. 17. */ht Fig. 4-19. Centers of pressure in plane sliders. Numbers refer to L/B ratio. As explained in Chap. 3, the results obtained for fixed sliders can be extended to cover the performance of pivoted-shoe thrust bearings. From the preceding analysis, the center of pressure can be written as 1 6/xC/L jl_^n-lna + a — 1 Wa2 [2 a2 - 1 1 (L/B)a(a - l)2 (ir/2)anynJi(anXi)[Ji(anz2) — aJ\{otnXi)] tanh anX2(a — 1)(L/2R) (<xnx2)z[Ji2(anx2) - Ji2(anxi)] (4-70) which relates the location of the pivot from the trailing edge to the 0.18 Q22 0.26 Q30 0.34 0.38 0.42 0.46 Q50 x/B x/B Fio. 4-20. The variation of angle of incli- Fig. 4-21. Friction coefficient as a func- nation with pivot position measured from tion of pivot position measured from trail- trailing edge. Numbers refer to L/B ing edge. Numbers refer to L/B ratio, ratio.
Incompressible Lubrication; Finite Bearings 129 0.44 * 0.36 0.28 > Q20 5: 0.12 0.04 0.14 0.22 0.30 0.38 0.46 i/B Fici. 4-22. Minimum film thickness as a function of pivot position measured from trailing edge. Numbers refer to L/B ratio. remaining bearing parameters. Figures 4-20 to 4-22 give the perform¬ ance of pivoted-shoe sliders. 4-13. Sector Pad; Computer Solutions. The solution of a sectorial thrust-bearing element by finite difference equations can be obtained by methods similar to those used on journal bearings. However, since the underlying differential equation and the geometry of the pad are differ¬ ent, new dimensionless ratios must be used. Referring to Fig. 4-23, we write f = i- U = 2-irrN = 2-rRifU til where h — hi — h2. The value of p — (p/nN)(8/L)2 is a dimensionless load factor for thrust bearings and can be considered the equivalent of the Sommerfeld number
130 Theory of Hydrodynamic Lubrication for journal bearings. By writing these substitutions in Eq. (1-9), we have <«■> By expressing the derivatives in finite difference form and using them in Eq. (4-71), we have for any point in the grid, 10—p (^\ h*,i+H I £ 3 (rp)<+hj I LJ 1 \L) Ad + ~(W' + '-*■* ~WF i zTs P*.j+1 i £ 3 Pt.i-i , _ + h‘-»» ¥~Ke + T~m Pi,j — : _L hj-iiJi-U i 1 (£3 . J. ^3 \ (Af)2 + (Af)2 + fw Aff 1 *J-M + (4-72) and this equation too is of the form Pi,j = do + CllPi+\.j 0>2Pi-\,j + ttsPtJ+l 4" 0>\Pi,j-\ and for a set of m X n equations, can be solved by methods previously described. The film shape assumed for the following solution is that of an uniform taper in the circumferential direction given by or in dimensionless form h = A, + t (l - (4-73) * “ 7 + (4 “ |) The expression for the load capacity is nt n =/r iopr d$ dr=low**** Af ^22 ^ (4-74) The dissipated power is *hN*R2a \n2/ j=1 i-i where j is the term in the brackets. <4-,s)
Inco?npressible Lubrication; Finite Bearings 131 The flow of lubricant is again defined in terms of the flow coefficient q, namely, irNLht (4-76) (/?i + Ri) + qnrR2NL6 (4-77) Ad Ri.Rt Af (4-78) (4-79) From above the two flow coefficients q, and qi are j - = _L (k\ qi 12t \Rt) A f de i Thus the final answers will be given as The results of the calculations are given in Table 4-7, and some of the relationships are plotted in Figs. 4-24 to 4-26. When the quantity Fig. 4-24. Ix)ad capacity of tapered-land thrust bearings. (iiN/P)(L/8)2 is plotted vs. the angular span 0, there is a minimum in the curve indicating the optimum number of pads for a given set of condi¬ tions, and these are tabulated in Table 4-8. Table 4-9 gives a comparison of results for an inclined slider, a slider with exponential film, and for a sector with a circumferential taper only. Again the closeness of the results emphasizes the conclusion that, once h\/h2 is fixed, the actual film shape is of secondary importance.
132 Theory of Hydrodynamic Lubrication Table 4-7. Performance of Sector Thrust Bearings L/Rt hi/6 ft deg U Q\n Center of pressure * j At Ri At Ri 6 r H 1 80 1.423 0.34 0.46 0.87 0.64 0.37 2.44 55 1.108 0.32 0.44 0.84 0.625 0.45 1.685 40 0.947 0.28 0.395 0.81 0.61 0.49 1.26 30 0.870 0.235 0.35 0.75 0.605 0.51 0.95 H 80 0.321 0.35 0.47 0.87 0.71 0.37 3.94 55 0.257 0.32 0.44 0.84 0.69 0.47 2.70 40 0.225 0.28 0.40 0.79 0.67 0.50 2.00 30 0.211 0.24 0.36 0.74 0.66 0.51 1.57 H 80 0.0855 0.35 0.47 0.87 0.78 0.41 5.96 55 0.0714 0.32 0.44 0.83 0.76 0.45 4.25 40 0.0652 0.29 0.41 0.78 0.74 0.505 3.23 30 0.0635 0.245 0.36 0.70 0.73 0.52 2.54 H 80 0.0278 0.36 0.48 0.85 0.83 0.465 8.51 55 0.0247 0.33 0.45 0.81 0.815 0.50 6.23 40 0.0238 0.29 0.41 0.75 0.795 0.51 4.86 30 0.0242 0.25 0.37 0.67 0.78 0.565 3.91 'A l 80 1.72 0.23 0.405 0.75 0.62 0.48 2.90 55 1.494 0.19 0.36 0.69 0.61 0.51 1.96 40 1.435 0.145 0.31 0.61 0.60 0.53 1.47 30 1.489 0.11 0.26 0.57 0.59 0.55 1.13 A 80 0.402 0.23 0.41 0.74 0.685 0.46 4.72 55 0.3585 0.19 0.33 0.61 0.67 0.52 3.33 40 0.352 0.15 0.31 0.60 0.655 0.53 2.49 30 0.370 0.11 0.26 0.53 0.65 0.55 1.92 A 80 0.1138 0.24 0.42 0.72 0.755 0.48 7.32 55 0.1062 0.20 0.27 0.65 0.735 0.52 5.29 40 0.1080 0.15 0.32 0.56 0.72 0.54 4.065 30 0.1103 0.11 0.27 0.49 0.71 0.56 3.18 % 80 0.0402 0.25 0.42 0.70 0.81 0.50 10.81 55 0.0399 0.20 0.28 0.62 0.78 0.53 8.06 40 0.0423 0.16 0.32 0.53 0.77 0.55 6.30 30 0.0470 0.11 0.27 0.44 0.765 0.57 5.01 n l 80 2.240 0.12 0.35 0.60 0.61 0.50 3.06 55 2.185 0.082 0.295 0.53 0.60 0.55 2.12 40 2.320 0.052 0.245 0.48 0.59 0.58 1.57 30 2.590 0.033 0.200 0.44 0.59 0.61 1.20
Incompressible Lubrication; Finite Bearings 133 Table 4-7. Performance of Sector Thrust Bearings (Continued) L/Ri ht/h 0, deg Qia Center of pressure* j At Rx At Ri 6 m r M 80 0.538 0.13 0.35 0.58 0.67 0.51 5.07 55 0.537 0.084 0.30 0.51 0.66 0.56 3.59 40 0.578 0.0535 0.25 0.45 0.65 0.59 2.70 30 0.653 0.034 0.20 0.40 0.645 0.61 2.07 H 80 0.1598 0.13 0.36 0.56 0.735 0.53 8.00 55 0.1655 0.087 0.30 0.46 0.72 0.57 5.79 40 0.1820 0.055 0.25 0.40 0.71 0.60 4.43 30 0.2085 0.035 0.21 0.36 0.705 0.62 3.46 H 80 0.0599 0.14 0.365 0.53 0.79 0.55 12.07 55 0.0649 0.09 0.31 0.44 0.78 0.58 8.98 40 0.0737 0.056 0.25 0.35 0.765 0.61 6.94 30 0.0861 0.036 0.21 0.29 0.76 0.63 5.47 0 = 6/0 r = (r — R\)/L. Fig. 4-25 bearings. Angular span 0, deg Side flow in tapered-land thrust 40 60 80 100 Angular span 0, deg Fig. 4-2G. Friction in tapered-land thrust bearings.
134 Theory of Hydrodynamic Lubrication Table 4-8. Optimum Geometry in Tapered-land Thrust Bearing l/r2 hi/8 0, deg Number of pads Vs 1 <30 >10 A <30 >10 H 35 9 A 40 8 A l 40 8 A 45 7 H 50 6 A 60 5 H 1 50 6 A 60 5 A 80 4 A >80 4 Table 4-9. Performance of Thrust Bearings with Various Film Configurations a Plane slider* Exponential sliderf Sector padf p _ PL'hS tiojRi4 2.00 0.0810 0.0819 0.0826 2.50 0.113 0.1137 0.106 2.85 0.135 0.135 0.125 Fhn F — ■ - IXOiRi* 2.00 0.66 0.81 0.78 2.50 0.74 0.875 0.825 3.04 0.84 0.95 0.88 * h = ax t h = kxek** J h = hi + 5(1 - 0/0) SOURCES 1. Tao, L. N.: On Journal Bearings of Finite Length with Variable Viscosity, J. Appl. Mechanics, vol. 26, June, 1959. 2. Fedor, J. V.: A Sommerfeld Solution for Finite Journal Bearings with Circum¬ ferential Grooves, Trans. ASME, Series D, June, 1960. 3. Sassenfeld, H., and A. Walter: Journal Bearing Calculations, VDl-Forschungs- heft (B), vol. 20, no. 441, 1954. 4. Kingsbury, A.: On Problems in the Theory of Fluid Film Lubrication with an Experimental Method of Solution, ASME Paper APM-53-5, 1930.
Incompressible Lubrication; Finite Bearings 135 5. Needs, S. J.: Effects of Side Leakage in 120° Centrally Supported Journal Bear¬ ings, Trans. ASME, vol. 56, 721, 1934. 6. Pinkus, O.: Analysis of Elliptical Bearings, Trans. ASME, vol. 78, July, 1956. 7. Sternlicht, B., and F. J. Maginnis: Application of Digital Computers to Bearing Design, Trans. ASME, vol. 79, October, 1957. 8. Pinkus, O.: Solution of Reynolds Equation for Finite Journal Bearings, Trans. ASME, vol. 80, May, 1958. 9. Raimondi, A. A., and J. Boyd: A Solution for the Finite Journal Bearing and Its Application to Analysis and Design—III, Trans. ASLE, vol. 1, no. 1, 1958. 10. Raimondi, A. A.: A Theoretical Study of the Effect of Offset Loads on the Per¬ formance of a 120° Partial Bearing, Trans. ASLE, vol. 2, no. 1, 1959. 11. Pinkus, O.: Analysis and Characteristics of the Three-lobe Bearing, Trans. ASME, Ser. D, vol. 79, March, 1959. 12. Pinkus, O.: Analysis of Journal Bearings with Arbitrary Load Vector, Trans. ASME, vol. 79, August, 1957. 13. Michell, A. G. M.: The Lubrication of Plane Surfaces, Z. Math. u. Physik, vol. 132, p. 123, 1905. 14. Archibald, F. R.: A Simple Hydrodynamic Thrust Bearing, Trans. ASME, vol. 72, May, 1950. 15. Charnes, A., and E. Saibel: On the Solution of the Reynolds Equation for Slider-bearing Lubrication: The Rectangular Thrust Bearing, Trans. ASME, 1952, vol. 74, p. 867. 16. Muskat, M., F. Morgan, and M. W. Meres: The Lubrication of Plane Sliders, J. Appl. Phys., vol. 11, March, 1940. 17. Hays, D. I.: Plane Sliders of Finite Width, Trans. ASLE, vol. 1, no. 2, 1958. 18. Pinkus, O.: Solution of the Tapered-land Sector Thrust Bearing, Trans. ASME, vol. 80, October, 1958. 19. Pinkus, O.: Solution of Reynolds Equation for Arbitrarily Loaded Journal Bearings, Joint ASME-ASLE Lubrication Conf., October, 1960.
CHAPTER 5 HYDRODYNAMIC GAS BEARINGS GENERAL CONSIDERATIONS Hydrodynamic gas bearings operate on essentially the same principles as the bearings described in preceding chapters. However, since the lubricant used is a compressible fluid, a number of divergencies from incompressible lubrication are introduced. These differences can be summarized as follows: 1. Pressure Distribution. In an incompressible fluid, a full converging- diverging film will have pressures greater and less than ambient. How¬ ever, this need not be the case for a compressible film, where the pressure may always be greater than ambient. The hydrodynamic pressures of an incompressible film are independent of the ambient pressure. Conse¬ quently, the absolute pressures can be determined by summing the pres¬ sure rise and the ambient pressure. This cannot be done for a compressible film. The absolute pressure must be used in the Reynolds compress¬ ible equation. The ability of a compressible film to carry a load increases with the ambient pressure. 2. Variation in Density. Unless the density is treated as a variable, the results will differ materially from actual bearing performance. Table 5-1 is an illustration of the kind and degree of such discrepancies. Table 5-1. Effect of Compressibility on Bearing Characteristics N, rpm U, ips € 230 72.2 0.39 805 252.5 0.1625 1,730 542.5 0.091 Value Exp. Theor. Exp. Theor. Exp. Theor. Vo - pa, psi 77.7 80.5 43.7 100.0 43.7 117.2 Pmin Pa, psi - 59.0 - 80.5 - 47.2 -100 - 50.7 -117.2 VO Praia, psi 136.7 161.0 90.9 200.0 94.4 234.4 6o — 180, deg - 28 - 57 - 43 - 76 - 43 - 82 0inin 180, deg.. . . 51 57 96 76 129 82 0.n»D — 00, deg 79 114 139 152 172 164 136
Hydrodynamic Gas Bearings 137 The theoretical results in Table 5-1 are based on air as an incompressible fluid at a constant atmospheric density. Several points emerge. The experimental data indicate that, as the velocity increases, the value of Po — Pa decreases and approaches a definite limit, that the position of the maximum pressure is displaced nearer to the point of closest approach, and that the position of minimum pressure is displaced further from that point. Only when the bearing number A approaches zero can incompress¬ ible theory be used to predict the behavior of gas bearings. 3. Viscosity. In general, the approach to the analysis of gas bearings is to assume an isothermal path for the lubricant. This is not too bad an assumption for moderate speeds, at which the frictional losses are small and there is only a slight temperature rise. In compressible fluids, unlike incompressible fluids, the viscosity increases with temperature. 4. Striation and Slip. There is no striation in gas bearings, and this simplifies considerably the boundary conditions in the analysis. When the bearing film thickness becomes comparable in magnitude to the molecular mean free path of the gas, continuum flow theory no longer holds and “slip flow” occurs. In order to bring the predictions based on the continuum theory and experimental results into better agreement, the Reynolds equation must be modified to include slip velocities at the boundaries. Perturbation methods may then be employed in the analysis. Examples of cases in which the influence of the molecular mean free path should be taken into consideration are hydrodynamic gas bearings operating with very small film thicknesses or at low ambi¬ ent pressures. Analysis indicates that, under these conditions, the load-carrying capacity decreases as the ratio of the molecular mean free path to film thickness increases. This decrease is most pronounced at low speed. For more detail, see Ref. 1. 5. Dimensional Accuracy. This becomes a major item in the operation of gas bearings for two reasons. One is that the films in gas bearings are appreciably thinner than the films in incompressible lubrication, and the minimum film thickness may be of the same order of magnitude as the surface roughness of the journal and bearing. The other is that any ellipticity or waviness of the mating surfaces will cause the fluid to alternately expand and compress with resulting low and high pressures which distort the pressure profile and flow pattern. 6. Heat Transfer. Because of the drastically lower shear losses, the amount of heat transfer out of gas bearings is less than in liquid lubricants and becomes significant only at very high velocities. While the relative constancy of viscosity and the simpler boundary conditions tend to simplify the analyses of gas bearings, the introduction of variable density makes the mathematics considerably more difficult. In addition, even if perfect-gas relations are used for the fluid, there still
138 Theory of Hydrodynamic Lubrication remains the problem of choosing a thermodynamic path for the gas. The simplest path is one of constant temperature, but even then the solutions for gas bearings tend to be complex, for the differential equa¬ tions become nonlinear. LIMITING CHARACTERISTICS By using the perfect-gas relation in Eq. (1-11), the normalized Reynolds equation can be written as d (pllnhzdp\ , d (p1/nh3 dp\ _ A dpl,nh d£ \T“ d2/ d~z \/T” 31/ ~d£~ ( ' in which A = §naUB/h\ pa, P = p/pay h = h/h2) P — n/na, x = x/B, and z = z/B. In journal bearings h2 is replaced by the radial clearance C and the breadth B by radius R. Equation (5-1) may also be written as d rfraQ^+W")] £ ffr affi"*1”")] = n + 1 /dpl'nR\ dx[p dx J dz [_ p dz J n \ dx J and for isothermal conditions the above equation reduces to d_ A3 ap2\ d_ /h* df \ _ dph dx dx) + dz dz) dx ' Equation (5-2) is of the same form as Eq. (4-29) for incompressible lubrication. It is only necessary to relate p2 and 2ph of the compressible equation to p and h of the incompressible equation to establish similarity. Thus incompressible solutions can be modified, by proper substitution, to apply to compressible cases. When the bearing number approaches zero, good approximations of the pressure distribution and bearing load of a compressible film may be obtained from a solution which assumes the film to be incompressible. On the other hand, when the load number approaches infinity, the right- hand side of Eq. (5-1) becomes very large and the solution approaches that of an infinitely long bearing. It also follows that, for the pressures to remain finite under these conditions, it is necessary that d(Pllnh) dx Thus, the pressure distribution approaches =gy (5-3) in which a — h\/h2 is the film-thickness ratio where the film pressure is ambient.
Hydrodynamic Oas Bearings 139 For large bearing numbers, the load supported becomes less sensitive to velocity increases, and must approach W = = a» fl i-dx - 1 (5-4) P°BL J0 h" as an asymptotic limiting value for A —* oo. The frictional force may be expressed in normalized form as For large bearing numbers this force becomes a constant given by Fa-.= r^-dx (5-6) Jo h which means that the dimensional frictional force increases linearly with velocity and has the same value for both surfaces of the film. Therefore the coefficient of friction / = F/W must increase linearly with bearing number for a converging film where A —> ». For more detailed discus¬ sion, refer to Ref. 4. INFINITELY LONG SLIDER BEARINGS 6-1. Parallel Surface. In this case h — 1 and integration of Reynolds equation (5-1) with dp/dz = 0 yields v 1 piM_ e/p = am d£ (5-7) where C\ is a normalized constant of integration. If we assume the viscosity to remain constant, the above equation can be integrated again to yield the following relations: 1. If ^| < 0 and Ci > p1/n, we get ^ CV(n Hr i) i- 1 = -Ax + C2 (5-8 a) Note that the pressure will be discontinuous in the limit A - dp
140 Theory of Hydrodynamic Lubrication If the fluid film is isothermal, p + Ci ln |p — Ci| = Ax + CA (5-8c) The above constants can be evaluated by substituting the proper bound¬ ary conditions. However, as with the incompressible case, unless a pressure above ambient is imposed at some point within the film, the above equations will yield zero load capacity. Fig. 5-1. Isothermal pressure distributions for parallel sliders. Figure 5-1 shows an isothermal pressure distribution for a pressure ratio of four. The first of these figures has the inlet edge pressurized; the second has the trailing edge pressurized. Figure 5-2 gives a comparison between isothermal and adiabatic conditions for A = 3. The bearing load may be obtained from (5-9)
Hydrodynamic Gas Bearings 141 By substituting Eq. (5-7) in Eq. (5-9), we get * - * X h - *3^ (21 + '-T^)] - 1 (5-10) in which the + is for dp/dx < 0 (p2 < 1) and the - for dp/dx > 0 (p2 > 1). In the limiting condition A —» 0 the bearing load can be 0 ai 0.2 03 0.4 05 06 0.7 0.8 Q9 1.0 x = x/B Fig. 5-2. Isothermal and adiabatic pres¬ sure distribution for air-lubricated par¬ allel sliders. x=0 —+U x--B x~-B — U jr = 0 Fig. 5-3. Plane inclined sliders. obtained by integration of (dp/dx)^-o — Ci, which yields W -P*~ 1 For the other limit 3(p2 + 1) for $ 0 dx IPa-.. = 0 1Pa-« = ?2 - 1 f->° dx T- <° dx 5-2. Plane Inclined Slider. For a thrust bearing such as shown in Fig. 5-3 the Reynolds equation with isothermal flow becomes ^ + 6Uh = c1 y dh p where h = hi — ax. By employing a new dependent variable given by ph (a/y)Zdt dh (a/v)p -6Ut + Cx h The form of the integral depends on the roots of the denominator, or on the sign of yC\a — 9U2y2 — A2. Within the general range of U this quantity is positive, hence (5-11) £, we have (5-12)
142 Theory of Hydrodynamic Lubrication Now the boundary conditions are p = pa for x = 0, h = hi and for x = B h — h2. From these two boundary conditions we can determine C1 and C2. To find the value and position of maximum pressure, we notice that, when dp/dx = 0, 6 nUph = Ci Hence the position x0 of the maximum pressure is given by the equation hi - axo _ 1 , 2((V6t/)2 hi 2 ap^h/t/n - GUpahi + Ci xo having been found, the value of the maximum pressure can be obtained from Eq. (5-13). Comparison of these results with the incompressible solutions shows that compressibility forces the position of maximum pressure closer to the point of minimum film thickness and decreases somewhat the value of the maximum pressure. Equations similar to Eqs. (5-13) and (5-14) are obtained for diverging inclined planes, except that tan-1 is replaced by a logarithmic form. Here the effect of compressibility is to displace the position of minimum pressure in the direction of motion and to reduce the differential between the generated subatmospheric pressures and the ambient pressure. Equation (5-13) may be somewhat easier to interpret if it is normalized. Under isothermal conditions, the viscosity may be considered constant and the equation describing pressure distribution may then be written as where Ci is a new integration constant, and it represents the product of pressure and the film thickness at the point where the maximum pressure occurs. The integration of Eq. (5-15) is simplified by expressing h as the d d independent variable. Since (a — 1) = — —> the above equation may be written as As before we choose a new dependent variable £ = ph. For the boundary conditions we set p(0) = 1 and p(l) = p2. Therefore, at the leading edge x = 0, £ = a, while at the trailing edge x = 1, £ = p2. For the plane slider bearing, p2 = I. The same methods apply for p2 7* 1. tan_1 aCi/6U - ZiiU _ ~ 3mIT ) (5-14) phs || = A {ph - Ci) (5-15)
Hydrodynamic Gas Bearings 143 Now, setting A/a — 1 = X, we get {({« - Af + ACi)"1 di = ^ h Integration is immediately possible, and it yields e - A£ + ACi = (5-16) in which C2 is a constant of integration and ^(£) assumes values as follows: ^i(£) = exp- (— -— tan-1 -—— 4—) for A < 4C\ l[A(4C»-A)]* [A(4Ci - A)]»J ^c(f) = exp 2 for Ac = 4Ci l2f - A - [A(4C, - A)]*] The load-carrying capacity may be obtained by integration of Eq. (5-16). W = (a- I)"1 W = (a- I)-1 A ln a + (A - Ci) lnj^^yjj for A < 4Ci Ac ln ~~ ~ 2 p2 — Ac - Ac)-1] J W = (a — I)-1 |a ln a + C, ln jj for A > 4C: - ^ [(2a - Ac)-1 - (2p2 - Ae)-1]} for Ae = 4C, Table 5-2 gives the isothermal load capacity and center of pressure for infinitely long slider bearings. For bearing numbers approaching zero the load-carrying capacity may be expressed as Wa-+ o = A (a2 + l)-»[ln a - 2 (a - 1 )(a + I)"1] - 1 For large bearing numbers with adiabatic conditions the bearing load is given by q(qn~1 — 1) (n - l)(a - 1) For isothermal conditions and Fig. 5-4 compares an isothermal and adiabatic pressure distribution for an inclined slider of infinite width operating with a load number equal to 400.
144 Theory of Hydrodynamic Lubrication Table 5-2. Performance of Infinitely Long Sliders under Isothermal Conditions* a A W/LBp. X 1.5 0.5 0.01091 0.5456 1.0 0.02172 0.552 5.0 0.0957 0.5861 10.0 0.1486 0.6168 25.0 0.1942 0.6585 2.0 0.5 0.01323 0.5724 1.0 0.02640 0.5761 5.0 0.1234 0.6008 10.0 0.2124 0.6235 30.0 0.3367 0.6705 50.0 0.3639 0.6870 3.0 0.5 0.01232 0.6095 1.0 0.02063 0.6116 5.0 0.1201 0.6264 10.0 0.2252 0.6402 50.0 0.5618 0.6950 4.0 0.5 0.01034 0.6349 1.0 0.02068 0.6364 5.0 0.1023 0.6464 10.0 0.1980 0.6560 25.0 0.4264 0.6756 50.0 0.6424 0.6976 6.0 0.5 0.007241 0.6688 1.0 0.01448 0.6696 5.0 0.07243 0.6753 10.0 0.1435 0.6809 50.0 0.5943 0.7057 100.0 0.9021 0.7255 * After W. A. Gross, IBM Corporation. ii a> Fig. 5-4. Isothermal and adiabatic pressure distribution for air-lubricated plane inclined slider. i 1 — a = 9 A A = 400 ■7 f n - 1 /l:U / / f\ J v r / * / V ✓ 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 x - x/B
Hydrodynamic Gas Bearings 145 The first moment may be obtained by evaluation of Eq. (5-16). For limiting conditions, the center of pressure may be expressed as f = Co - i -1-6 a(a + 2) ln a - (5a - l)(a - 1) A~° 1 ' 2[(a2 - 1) ln a - 4(a - l)2] When the bearing number approaches infinity and the gas path is poly¬ tropic, the center of pressure may be expressed by ^ = (a - 1)- a - a^-)(n -2)-V + H(l-a-)] v ' I (1 - al~n)(n - l)_lan + 1 - a J If the film is isothermal, the center of pressure becomes £a-.oo = a(a - I)"1 - (a - l){2[a(ln a - 1) + I])”1 Figure 5-5 shows the effects of bearing number and film-thickness ratio upon the center of pressure. Unlike the incompressible solutions, the Fig. 5-5. Center of pressure for isothermal infinitely long plane inclined sliders. center of pressure is determined completely not by the film thickness ratio a but by £ = /(a,A). For pivoted bearings, compressible analysis shows that a discontinuity may set in, in which a decreasing minimum film is accompanied by a decreased load capacity. In the case of a crowned slider, it is possible to have as many as two discontinuities, as shown in Fig. 5-6. To avoid this condition, it is first necessary to establish the range of load numbers to which the bearing is subjected and then pick £, making sure that for each a there will result an increase in W, as shown in Fig. 5-5 and Table 5-2. 5-3. Composite Slider. There are various combinations which may be considered composite slider bearings. One such configuration consisting
146 Theory of Hydrodynamic Lubrication Unstable \ region U B Fig. 5-6. Load vs. minimum film-thick- Fig. 5-7. Composite slider. ness instability. of an inclined slider and a flat is shown in Fig. 5-7. As discussed in Chap. 3, the bearing can be treated in two parts. Let The boundary conditions are p(xi = 0) = p(x2 = 1) = 1; also p(x i = 1) = p(x2 = 0) The interior pressure pe, as well as the necessary constants for an iso¬ thermal film, may be obtained by utilizing Eqs. (5-8c) and (5-16) Continuity of mass flow may be used as the fifth equation in order to solve for the five unknowns Note that the normalized pressure gradients will not be continuous unless Bi = B2. Equation (5-15) may be utilized to demonstrate that C\ = C3. The solution of Eqs. (5-17) is now possible, although very tedious. Equation (5-2) may be employed to gain an approximate idea of the limiting pressure distribution. By letting A —> 0 and neglecting z deriva¬ tives, the resulting pressure distributions are Ai = —^ A A2 = -77 A and Ai = 1— d a — 1 a — Aid + A1C1 = CM(P) pc2 — AiPc + A1C1 = Ctf(l) Pc + Ci ln \p. - C,| = Ci (5-17) 1 -|- Ci ln |1 — O3I = Ai C4 (5-18) Pa—»o = (pc2 - 1) (j2 - 1) (a2 - 1)-' + l]M Pa~»0 = [pc2 — (pc2 — 1)^2]^ 0 < Xi < 1 0 < X2 < 1
Hydrodynamic Gas Bearings B 147 0 0.2 0.4 0.6 0.8 —: X\/B{ |l 0.2 0.4 0.6 0.8 1.0 * 0 0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1.0 *~*iz*i!B\ Fig. 5-8. Isothermal pressure distribution in composite bearings. An additional relation is needed to evaluate pc, and this can be obtained from the pressure gradient relationship. dp dxi dp dx2 a2(p2 - 1) A_o h2(a + l)[h2 + (Pc - l)(a2 - h2)(a2 - l)~']» = - \ (Pc2 - l)[p«2 - (Pc2 - A—>0 * 0 < x\ < 1 0 < X2 < 1
148 Theory of Hydrodynamic Lubrication In the ca.se when A —► Eq. i.Voi may be applied. The pressure gradient is again discontinuous at the inner boundary. The pressure profile for two different film-thiekness ratios is shown in Fig. 5-8. Results are shown for length ratios B2 B\ = 0.5 and 1. The effect on pressure of altering the B+ 'Bi ratio for a fixed B is seen in Fig. 5-9 for which a = 2.33 and A = 26.6. The effect of bearing number upon isothermal load capacity and center of pressure is illustrated in Fig. 5-10 for a — 1.5, 2.33, 4, and 9. 5-4. Step Slider. The pressure distribution for the step slider may be obtained from Eqs. (5-8). By imposing the appropriate boundary con¬ ditions on Eq. (5-8c) we obtain the following system of four equations in five unknowns: i 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 / = x/B Fig. 5-9. Pressure distribution for various composite bearings. 1 — Cl In 11 — Cil = C2 pe -f Cl ln I pc - Ci| = Ai + C2 pc C3 ln | pc — C3I — Ci 1 -|- C3 ln 11 — C3I = Ao + Ci (5-19) where a — _ ilf — _ _ 1 B2 a2 B a2 B\ A2 B\ A and Since the continuity of mass must be preserved, (5-20) Thus from Eqs. (5-19) and (5-20) the four constants and the pressure at the step are evaluated. Note that in the leading region of the bearing,
Hydrodynamic Gas Bearings 149 the pressure gradient will monotonically increase, while in the trailing region it will monotonically decrease. Figure 5-11 shows the pressure distribution for several B2/Bi ratios. n i* ■o o * 2 WBS =9 / r4 1 ^ 1 1 - r—- <7=2.; <7=1.! V 05^. 1 / 0.5 0 n 0.8 iiw 0.7 £ 0.6 1.0 0.8 £ ? 0.6 * ^Q4 0 50 100 150 200 250 300 350 400 450 500 oo (<7) Bearing number A = 6\i0UBlpahf 0.2 */Js o . , 1 1 BJBS =0\J\ s 1^ ^0.5 <7 = 9 /// // A r ~r“r' r\ IO AfA' 0.8 0.7 «; 0.6 ? U) 5 10 15 20 25 30 35 40 45 50 Beoring number A= §\i0UBIp0h} Fig. 5-10. Isothermal bearing load and center of pressure for inclined-plane and composite bearings. 6-5. Convergent-Divergent Slider. Let us consider next a converging- diverging film as shown in Fig. 5-3. The same integral forms are obtained for sections AB and BC as in the separate cases [refer to Eq. (5-13)], but whereas the constant C\ had different values in these cases, it is easily seen that, owing to continuity, C\ must have the same value for a given velocity in both sections.
150 Theory of Hydrodynamic Lubrication Fig. 5-11. Isothermal pressure distribution for step bearings. Let pc be the pressure at B; then for the section AB we have the conditions p = p« at x = 0; p = pc at x = B. By substituting these in Eq. (5-13) and subtracting, we obtain 1 . apa2h\2/y — 6Upahi + C\ , h\ 2 ap2h22/m - 6Upch2 + Ci hi + W (tan- - tan- ap'hi ~ = 0 (5-21) where 42 = pCi« — 9U2y2. By treating the equation for p between B and C in the same way, we have 1 . apa2hi2/ti + QUpghi — Ci _ . hi 2 n otpc2ha2/n + QUpch2 — Ci n hi 3(JlU . (C2 + apa^l + 3/uC)(C2 — apehi — 3/xC) _ _ 00. + 2C; (C2 - apji! - 3mC)(C2 + ctpAi + ZvU) ~~ K } where C22 = pCia + 9 U2y.2.
Hydrodynamic Gas Bearings 151 Equations (5-21) and (5-22) determine the value of the constants Ci and pc. Once these are determined, it is possible to calculate the position and magnitude of maximum and minimum pressure as a function of velocity, plane inclination, and minimum film thickness. 6-6. Curved Slider. The film thickness in Fig. 5-12 may be ex¬ pressed with good accuracy by the parabolic approximation h = ha^ for 6 « 0 In case the surface y = 0 is also curved, the R in the above equation is replaced by an equivalent radius. It is convenient in dealing with films of this type to define a new angular variable £, such that x = (2hmiuR)H tan £ If now the dimensionless parameters * = (2 and * = h~ are employed, it follows that x = tan £ and h = sec2 £ By substitution of the above expressions into Eq. (5-15) and integra¬ tion, it follows that | = a(coS^-C'-C"S‘J) (5-23) where A = QpaU(2hminR)^/pahllia and U is the relative velocity at x — 0. Equation (5-23) for low bearing numbers may now be integrated. The results in terms of the transformed coordinate are ps.—+ o = 1 - ^ (2 sin 2£ + sin 4£) For large bearing numbers, Eq. (5-3) may be used. In this equation hi = hi/hnin is the film thickness ratio where the pressure is ambient. For ordinary convex slider bearing the pressure distribution may there¬ fore be written Fig. 5-12. Curved slider. pA—>oq = (a cos2 £)" = [a(l + ^2)“1]n
152 Theory of Hydrodynamic Lubrication FINITE SLIDER BEARINGS 6-7. Plane Inclined Slider. By setting the Reynolds equation in difference form, it is possible to solve it by numerical methods and obtain the pressure distribution in the x and z directions in a manner similar to that outlined in Chap. 4. Tables 5-3 and 5-4 give the dimensionless coefficients of load, center of pressure, and friction force for various ratios of inlet to outlet film thickness a, length-to-breadth ratio, and bearing numbers A for the case of plane inclined sliders under isothermal condi¬ tions. Using these coefficients it is possible to calculate such bearing characteristics as Load = W = W (where W = f 0 \ A 1 Friction force = F = BL jj r* dx dz = ^2fo^A F i jj (p — l)x dx dz Center of pressure £ = jj (p - 1) dx dz o INFINITELY LONG JOURNAL BEARINGS 6-8. Journal Bearing with Inertia Considered. By adding p Du/dt to Eq. (l-3a) we get the differential equation which governs steady-state flow without side leakage and with inertia effects considered, or d2u dp Du dp ( du du\ . = + p dt =Tx + l,{ua-x + vd-y) (5'24) The flow must satisfy the continuity equation which, for a compressible fluid, is from Eqs. (1-2) L(pu)+ty{pv) = 0 and which can be expanded to For a perfect gas du dp dv dp n dx dx dy dy
Hydrodynamic Gas Bearings 153 Table 5-3. W, Z, F Coefficients for Various Values of a and L/B A = 10 L/B = 0.25 A = 10 L/B - - 0.40 CL W X P w 2 F 1.0 0 0.50000 1.00000 0 0.50000 1.00000 1.5 0.00705 0.58987 0.81269 0.01532 0.59445 0.81476 1.8 0.00859 0.61923 0.73817 0.01873 0.61688 0.74222 2.0 0.00917 0.63585 0.69773 0.01996 0.62977 0.70313 2.2 0.00945 0.64749 0.66272 0.02071 0.64125 0.66947 2.5 0.00980 0.66634 0.61821 0.02126 0.65644 0.62680 2.8 0.00994 0.68309 0.58096 0.02137 0.66988 0.59124 3.0 0.00996 0.69311 0.55926 0.02128 0.67816 0.57058 4.0 0.00968 0.72726 0.47662 0.02004 0.70989 0.49216 5.0 0.00921 0.75199 0.42078 0.01827 0.73362 0.43890 6.0 0.00855 0.77289 0.37972 0.01655 0.75163 0.39974 7.0 0.00792 0.78869 0.34808 0.01502 0.76585 0.36938 A = 10 L/B 0.50 A = 10 L/B = • 0.75 1.0 0 Q.50000 1.00000 0 0.50000 1.00000 1.5 0.02150 0.59864 0.81630 0.03564 0.60591 0.81984 1.8 0.02623 0.61762 0.74523 0.04382 0.61928 0.75226 2.0 0.02792 0.62870 0.70711 0.04672 0.62742 0.71651 2.2 0.02891 0.63872 0.67440 0.04836 0.63489 0.68606 2.5 0.02958 0.65194 0.63305 0.04940 0.64496 0.64791 2.8 0.02965 0.66384 0.59869 0.04933 0.65444 0.61641 3.0 0.02945 0.67121 0.57875 0.04880 0.66042 0.59810 4.0 0.02722 0.70107 0.50293 0.04405 0.68539 0.52817 5.0 0.02447 0.72302 0.45130 0.03885 0.70432 0.48005 6 0 0.02195 0.73983 0.41323 0.03422 0.71944 0.44390 7.0 0.01975 0.75325 0.38357 0.03032 0.73174 0.41528 A = 10 L/B = = 1.00 A = 10 L/B = « 1.50 1.0 0 0.50000 1.00000 0 0.50000 1.00000 1.5 0.04630 0.60964 0.82250 0.05938 0.61272 0.82577 1.8 0.05768 0.62006 0.75780 0.07531 0.62044 0.76486 2.0 0.06172 0.62662 0.72401 0.08115 0.62551 0.73372 2.2 0.06401 0.63278 0.69546 0.08448 0.63045 0.70774 2.5 0.06545 0.64118 0.65995 0.08654 0.63734 0.67576 2.8 0.06532 0.64923 0.63080 0.08646 0.64399 0.64982 3.0 0.06452 0.65438 0.61382 0.08533 0.64840 0.63463 4.0 0.05768 0.67642 0.54862 0.07572 0.66775 0.57567 5.0 0.05034 0.69372 0.50305 0.06542 0.68353 0.53320 6.0 0.04393 0.70777 0.46817 0.05651 0.69663 0.49964 7.0 0.03856 0.71945 0.44001 0.04918 i 0.70764 0.47185
154 Theory of Hydrodynamic Lubrication Table 5-3. W, x, F Coefficients for Various Values of a and L/B (<Continued) A = 10 L/B = 2.00 A = 10 L/B = = 3.00 CL W X F W X F 1.0 0 0.50000 1.00000 0 0.50000 1.00000 1.5 0.06643 0.61390 0.82754 0.07354 0.61486 0.82932 1.8 0.08509 0.62045 0.76877 0.09504 0.62042 0.77275 2.0 0.09205 0.62489 0.73917 0.10321 0.62429 0.74475 2.2 0.09606 0.62929 0.71469 0.10796 0.62824 0.72182 2.5 0.09853 0.63556 0.68475 0.11087 0.63401 0.69401 2.8 0.09851 0.64163 0.66067 0.11094 0.63956 0.67186 3.0 0.09719 0.64574 0.64649 0.10946 0.64341 0.65877 4.0 0.08600 0.66397 0.59110 0.09662 0.66074 0.60702 5.0 0.07398 0.67910 0.55031 0.08279 0.67531 0.56794 6.0 0.06364 0.69175 0.51746 0.07095 0.68759 0.53572 7.0 0.05515 0.70247 0.48978 0.06126 0.69806 0.50811 A = i—* o \ to II 4.00 A = 10 L/B = = 5.00 1.0 0 0.50000 1.00000 0 0.50000 1.00000 1.5 0.07706 0.61527 0.83020 0.07911 0.61552 0.83071 1.8 0.09996 0.62040 0.77472 0.10287 0.62039 0.77588 2.0 0.10872 0.62404 0.74751 0.11198 0.62389 0.74914 2.2 0.11384 0.62779 0.72535 0.11732 0.62753 0.72744 2.5 0.11702 0.63332 0.69863 0.12065 0.63293 0.70135 2.8 0.11713 0.63867 0.67743 0.12079 0.63817 0.68072 3.0 0.11558 0.64240 0.66488 0.11920 0.64183 0.66850 4.0 0.10191 0.65932 0.61497 0.10505 0.65853 0.61968 5.0 0.08718 0.67365 0.57673 0.08979 0.67271 0.58194 6.0 0.07459 0.68576 0.54483 0.07675 0.68472 0.55022 7.0 0.06431 0.69610 0.51725 0.06611 0.69499 0.52266 A = 10 L/B = 10.00 A = 10 L/B « = 00 1.0 0 0.50000 1.00000 0 0.50000 1.00000 1.5 0.08301 0.61597 0.83168 0.08522 0.61625 0.83224 1.8 0.10835 0.62038 0.77807 0.11159 0.62037 0.77937 2.0 0.11818 0.62363 0.75224 0.12188 0.62345 0.75409 2.2 0.12395 0.62706 0.73142 0.12783 0.62678 0.73375 2.5 0.12758 0.63221 0.70655 0.13158 0.63178 0.70954 2.8 0.12779 0.63723 0.68703 0.13208 0.63659 0.69089 3.0 0.12614 0.64077 0.67544 0.13084 0.63979 0.68014 4.0 0.11105 0.65705 0.62867 0.11476 0.65605 0.63423 5.0 0.09476 0.67096 0.59187 0.09781 0.66979 0.59798 6.0 0.08086 0.68278 0.56049 0.08337 0.68148 0.56678 7.0 0.06954 0.69290 0.53294 0.07159 0.69156 0.53910
Hydrodynamic Gas Bearings 155 Table 5-4. W, x, P Coefficients for Various Values of a and L/B A = 0.01 L/B = 1.00 A = 0.01 L/B = 00 o W m X F W X F 1.0 0 0.50000 1.00000 0 0.50000 1.00000 1.5 0.04651 0.55401 0.82256 0.10912 0.54560 0.83821 1.8 0.06116 0.57199 0.75920 0.13881 0.56068 0.79026 2.0 0.06597 0.58310 0.72613 0.14821 0.56984 0.76725 2.2 0.06838 0.59336 0.69808 0.15222 0.57835 0.74838 2.5 0.06925 0.60731 0.66280 0.15206 0.58995 0.72490 2.8 0.06823 0.61966 0.63341 0.14790 0.60036 0.70512 3.0 0.06699 0.62727 0.61629 0.14400 0.60669 0.69330 4.0 0.06031 0.65446 0.55257 0.14868 0.61839 0.68511 5.0 0.05344 0.67453 0.50924 0.11263 0.64362 0.62762 6.0 0.04620 0.69213 0.47386 0.09134 0.66156 0.58670 7.0 0.03992 0.70737 0.44409 0.07630 0.67566 0.55322 A - 1.00 L/B = 1.00 A = 1.00 L/B = 00 1.0 0 0.50000 1.00000 0 0.50000 1.00000 1.5 0.05167 0.55815 0.82385 0.12007 0.55260 0.84095 1.8 0.06301 0.57675 0.75994 0.14245 0.56790 0.79171 2.0 0.06733 0.58758 0.72681 0.14976 0.57663 0.67803 2.2 0.06960 0.59733 0.69881 0.15326 0.58450 0.74900 2.5 0.07042 0.61045 0.66368 0.15304 0.59516 0.72564 2.8 0.06943 0.62212 0.63449 0.14892 0.60475 0.70604 3.0 0.06820 0.62920 0.61751 0.14505 0.61064 0.69436 4.0 0.06159 0.65470 0.55448 0.12531 0.63346 0.65006 5.0 0.05471 0.67369 0.51178 0.10517 0.65145 0.61269 6.0 0.04746 0.69021 0.47701 0.08905 0.66557 0.58098 7.0 0.04118 0.70434 0.44787 0.07668 0.67667 0.55434 A = ■ 40 L/B = = 1.00 A = 40 L/B = = 00 1.0 0 0.50000 1.00000 0 0.50000 1.00000 1.5 0.02253 0.66171 0.81656 0.02938 0.67002 0.81828 1.8 0.03119 0.66922 0.74721 0.04286 0.67580 0.75188 2.0 0.03550 0.67358 0.71090 0.05051 0.67865 0.71840 2.2 0.03886 0.67757 0.68037 0.05722 0.68086 0.69138 2.5 0.04246 0.68299 0.64270 0.06562 0.68324 0.66008 2.8 0.04469 0.68787 0.61223 0.07212 0.68486 0.63692 3.0 0.04566 0.69091 0.59497 0.07541 0.68572 0.62472 4.0 0.04630 0.70416 0.53155 0.08162 0.68999 0.58452 5.0 0.04319 0.71585 0.48875 0.07861 0.69502 0.55958 6.0 0.03931 0.72591 0.45662 0.07153 0.70146 0.53718 7.0 0.03549 0.73475 0.43078 0.06402 0.70800 0.51638
156 Theory of Hydrodynamic Lubrication Since T is constant and p varies with x only, we have ^ + + = 0 (5-25) dx dy p dx Equations (5-24) and (5-25) must now be solved for pressure. Chapter 12 describes in detail a method of averaging the inertia of the fluid across the film height by writing / du , du\ p fh ( du . du\ »{UTx + VlTy) = hjo {UTx + VlTy)dy We thus have by this approach d2U 1 dp p fh ( du du\ j \ /c o W = vTx + lTh]o \uTx + vTy)dy=Hx) (5-26) where after the integration is performed the right-hand side is a function of x only. Integrating Eq. (5-24) twice and evaluating the constants of integration by the boundary conditions w(0) = U and u(h) = 0 yields u = f(x) By substituting this u into Eq. (5-25) and integrating, we obtain „ = (M- E\?£dh _ (y* _ hy*\ df(x) \ 2 h2) 2 dx \3 2 / 2dx where Ci is an integration constant which by the condition that v(0) = 0 must be zero. Now by substituting u and v into Eq. (5-26) we obtain an expression for f(x) in terms of its derivative with respect to x. Another such equation can be obtained by applying the boundary condition v(h) = 0 in the expression for v. The derivative oif(x) can be eliminated between these two equations, leaving a single equation in f(x). Elimination of f(x) between this equation and ^ [h , [Uh f(x)h*l J0 ^ = p ^ "2 12~ J where G' is the flow per unit length, leads to the expression 1 dp _ QU 12& = 2p ["9Gft dh y. dx h2 ph3 \5y(RTg [p2h3 dx
Hydrodynamic Gas Bearings 157 Upon writing for the film thickness C(1 + « cos 6) and using the substi¬ tutions - = JL (9X A = G'C(S{T r = R(a v mw\^/ nR3o)2 y/g&T the expression for the pressure gradient becomes dp 6 12 A _ 2B2 r pe sin d dd (1 + € cos 6)2 p(l € cos 0)3 15 |_1 + € cos 6 9A2tsin9 dp( 9A2 18 X] (5 2S) p(i + e cos ey T de y T ps(i +«cos ey 1 +«cos e)\ K J The quantity A in this equation, and the constant which will arise when the equation is integrated for pressure, can be evaluated by the pressures specified at the beginning and the end of the bearing arc. Equation (5-28) is difficult to integrate exactly, but it may be solved numerically by using the Runge-Kutta method of integration. This gives rise to a trial-and-error solution wherein the pressure is set equal to atmospheric at one end of the film (this evaluates the integration con¬ stant) and the value of A is so adjusted that the pressure reduces to atmospheric at the other end. 5-9. Journal Bearing with Inertia Neglected. The right-hand side of Eq. (5-28) is a measure of the lubricant inertia. By neglecting these terms and replacing p and A by their dimensional parameters, we obtain dy _ §nUR I"i Ci 1 /c oq') de c2(l + e cos e)2 [ C(i + c cos $)p J y} This equation can also be obtained directly from Eq. (1-11) by setting dp/dz = Vo = 0 and using the isothermal relation p « p. Equation (5-29) can be integrated numerically by using the Runge-Kutta method. The bearing arc is divided into a number of segments. To begin the evaluation, two constants are assumed. These two constants for prac¬ tical purposes are the pressure and its gradient at the beginning of the first segment. The test of the correctness of these assumed values is (1) the pressure at d = 0 must be equal to that at 6 = 2t and (2) from con¬ tinuity the total mass of air must be constant or f** (p - Pa) (1 + cos 6) dd = 0 The assumed values of the constants have to be varied until these condi¬ tions are satisfied. At high speeds the inertia effect can be significant for gas-lubricated journal bearings. Table 5-5 compares some results obtained from both
158 Theory of Hydrodynamic Lubrication Eqs. (5-28) and (5-29) and shows the effects of inertia on the operation of a partial 180° gas-lubricated journal bearing in laminar flow. (Transi¬ tion occurs at Ne = 300,000 rpm.) Table 5-5. Effect of Lubricant Inertia on Performance of Journal Bearings e Speed, rpm W/L 0 Inertia considered Inertia neglected Inertia considered Inertia neglected 0.2 25,000 6.65 6.65 32.3 32.0 50,000 8.33 8.65 27.2 26.5 100,000 8.66 9.15 25.7 22.0 150,000 8.66 9.58 26.3 . 20.8 200,000 8.66 9.60 28.0 20.3 0.4 25,000 16.6 16.7 28.0 27.1 50,000 20.0 20.8 24.0 21.8 100,000 21.6 22.9 23.1 18.4 150,000 21.6 23.3 24.5 18.0 200,000 21.2 23.5 26.8 17.5 0.6 25,000 33.3 33.4 23.2 22.0 50,000 40.0 41.7 20.7 17.5 100,000 44.1 46.2 21.2 15.3 150,000 44.3 47.9 23.5 14.5 200,000 43.3 48.8 26.0 14.0 D = 2 in., C/R = 0.001. At 6i = 45° and 0a = 225°, p = pa. 0i = angle between line of centers and inlet edge. 5-10. Numerical Solution. Harrison8* in his solution assumed a dis¬ continuous linear film variation in which the mean film thickness is the same as the radial clearance of the journal bearing. He was therefore able to use the equations for a slider bearing in the journal bearing analysis. He further assumed the flow to be isothermal. Sheinberg7 applied numerical methods directly to the pressure gradient Table (5-6) shows results of his computation for attitude angle, load, and frictional moment. * Such superscript figures indicate references listed under Sources at the end of the chapter.
Hydrodynamic Gas Bearings 159 Table 5-6. Performance of Infinitely Long Journal Bearings A (C/Amin)*A t <t> W* Af* 0 0 0.333 90.0 0 7.7 0.334 0.498 0.333 68.0 0.332 7.5 0.542 1.224 0.333 58.6 0.518 7.4 0.858 1.932 0.333 49.6 0.712 7.3 1.074 2.412 0.333 41.6 0.798 7.2 3.336 7.50 0.333 17.6 1.312 6.8 00 00 0.333 0 1.52 6.7 0 0 0.500 90.0 0 9.7 0.125 0.498 0.500 78.6 0.244 9.5 0.780 3.12 0.500 47.6 1.04 8.7 1.332 5.34 0.500 36.6 1.51 8.3 2.358 9.426 0.500 25.0 2.06 7.9 5.004 20.04 0.500 13.2 2.58 7.5 oo 00 0.500 0 2.92 7.3 0 0 0.625 90.0 0 12.0 0.936 6.648 0.625 42.0 1.65 10.1 1.722 12.24 0.625 29.3 2.48 9.3 3.168 22.56 0.625 19.3 3.44 8.7 oo 00 0.625 0 4.59 8.0 0 0 0.714 90.0 0 17.2 1.008 12.36 0.714 35.3 2.15 13.8 2.016 24.6 0.714 25.6 3.63 12.5 3.810 46.68 0.714 16.0 4.94 11.7 oo 00 0.714 0 6.46 10.7 * W = W/VaRL M = MChaUR'L 6-11. Katto and Soda Solution. By transforming the variables p and 0 into ^ and 0 and using series expansion, it is possible to solve Eq. (5-29). This solution is reasonably accurate even for large eccentricity ratios. The conditions which the solution of Eq. (5-29) must satisfy are as follows: 1. The pressure p must be a continuous and periodic function of 0, its period being 2tt. 2. The total mass of air contained in the bearing clearance must be constant or ph dd = 2irCpa (5-30) where pa is the pressure which determines the mass of gas in the bearing clearance. Now we take the notations A, s, r, and v and transform the function p and the variable 0 into ^ and /3 as follows:
160 Theory of Hydrodynamic Lubrication A = 6 nU s2 = 1 - €2 r = 1 — e cos 0 where r is a function of 0 and v is a constant which is used instead of C\. By using these notations, we obtain from Eq. (5-29) 3-;(-*) ««» In this equation \p also must be a continuous and periodic function of 0 with a period of 2*\ If we consider only the case of v > 0, it is clear from Eq. (5-31) that the sign of d\p/d0 is determined by r < ^, since r cannot be negative for any value of 0 (e can take only the values between 0 and 1). Accordingly, we can state that the solution of Eq. (5-31) as a continuous and periodic function of 0 must exist in the region 1 + * ^ ^ 1 - € Next, we consider the solutions for the particular cases when the constant v becomes infinitely small or large. In the former case, since dp/d0 must be finite for any value of 0y we obtain from Eq. (5-31) lim„_»o yp = t In the latter case, as v increases to an infinite value, d\p/d0 approaches zero. By multiplying Eq. (5-31) by \J/n, (— oo < n < oo), we have ypn d\p = (T\pn — Thpn~l) d0 (5-32) and the integral of the left-hand member of this equation between 0 = 0 and 2ir is zero, since \p is a periodic function of 0. We have therefore Jq2v Tipn d0 = p rVn-1 d0 When we take \p as a constant in this formula, we obtain the following solution €2 lim^oo \p = 1+2 Considering that r/\p is always positive, we obtain cos 0 = CC i * AR € -f~ COS d 1 -h € cos 6 * v i(£)
Hydrodynamic Gas Bearings From this, combined with Eq. (5-31), we have 161 i*k < i rd& Now, by rewriting Eq. (5-31), we have , = 1 * T1 - (v/r)(dtm which can be expanded as follows: By disregarding all terms of the second or higher order, we obtain an approximate solution * = 1 - i-qjrjicoe|8 + sin 0 (5-33) The validity of this approximation decreases with an increase in v and e. However, the magnitude of v appearing in bearing problems is not very large, so that the error caused by the above approximation is small. By transforming p, 0, etc. of Eq. (5-30) to our notation, we have ■-j ycWM) (5-34) 0 « v /. Now, by multiplying Eq. (5-32) by r_m (m > 0), we have 7. _ 1 / V _ ^-1\ Tm W v ^rm-l Tm- 2 J From the transformation between p and \f/ and the magnitude of r it is seen that the integral of the left-hand member is positive between p = 0 and 2r. Therefore we obtain [2w jtL ^ f2’ t!ll JR Jo rm_1 Jo r"-2 and if n = 1, m = 3 in this inequality, we have r2wj, f2w 1 / —2dp> -dp (5-35) Jo T Jo r Moreover, the allowable maximum value for ^ is 1 + e and we have a + >>/„”? >£*•«’ <M6) Thus we have finally from Eqs. (5-34) to (5-36) the following inequality:
162 Theory of Hydrodynamic Lubrication from which it is seen that the magnitude of v appearing in bearing prob¬ lems is of the same order as s{C2pJ AR). Both these values are small and may be neglected. Now by converting ^ and 0 in Eq. (5-33) to the original notations, we can obtain the pressure distribution p as a function of 0, involving Coefficient of friction (R/C)f e and v as parameters. The relation between the velocity of the bearing A — GpU and the parameters e and v is obtained by sub¬ stituting the above pressure into Eq. (5-30). The integral of the above film pressure must balance the bearing load in both magnitude and direction. From these conditions we can obtain the relations between the attitude angle <f>, the mean load P, the velocity of the journal, and parameters e and v. The moment or the friction coefficient can be easily calculated in the same way as for the usual bearing problem.
Hydrodynamic Gas Bearings 163 In order to represent the results in dimensionless form, the following parameters are used: A — ^ ^ ^ - — P ~ C2Pa ~ Pa’ P ~ Fa By using the above notations, we obtain 1 I „2 o2 A = 2 (5-38) vs2-v2 (5-39) tan <f> = v (5-40) _ _ 1 + v2 [" € / e + cos 6 \ tsv / sin 0 \1 V ~ s2 + V2 [ 1 + »! \1 + e c°s e) + 1 + r \1 + e cos 0/ J (5-41) / _ i * Vi + v2 r« , /. _n1 (za<)'i C/R 21-s v [3*1 + *^ ( ^ In Eq. (5-42) the positive and negative signs represent the coefficient of friction for the journal and bearing, respectively. The various bearing solutions are shown in Fig. 5-13. FINITE JOURNAL BEARINGS 6-12. Perturbation Solution. Another approach to the solution of gas- lubricated journal bearings is to use perturbation methods to replace the nonlinear Reynolds equations by an infinite set of linear differential equations. A sufficient number of these equations are then solved. This method, like the other gas lubrication approximations, has greatest accu¬ racy for small eccentricities. The value of the solution lies in its com¬ paratively simple form, which enables us to handle general processes for the lubricating film. The two-dimensional Reynolds equation for constant loading with density proportional to plln is from Eq. (1-11) d_ dx y ■ 2) + i (r'lnh3 S) = I W'V {5-43) By use of x = R6} z = z/R} and division by pa1/nC3, Eq. (5-43) becomes & [©“" ©’ S]+ h [(£)'" ©’ If] = *** i [©“" &] (5-44) where A = - — ( npa\CJ
164 Theory of Hydrodynamic Lubrication The partial differential equation (5-44) is nonlinear in p, but a perturba¬ tion solution can be developed if pressure is expressed as a power series of increasing powers of e: V = Po + «Pl -f €2p2 -f €3P3 + • • • By writing h in terms of 6 and using p from above in (5-44), the follow¬ ing set of linear partial differential equations is obtained W - A + S' " ~lnr' "" * < W8*) S - * %'+S - "»i [(&)'+(£)’] + ( 3 sin 6 - 2A cos 6 - A ^\ ^ (5-45b) \ Pa/ O0 <?Pl - + d2Pl = . . . (5-45^ det d$ + dz'1 (5-45c) The first equation can be solved for pi, which can then be substituted into the right-hand side of the second equation as a known function. This permits the second equation to be solved for p2, which in turn can be used to obtain the solution for p3. Because the “driving functions” on the right-hand side soon become quite lengthy, the process is, in prac¬ tice, restricted by algebraic complexity to obtaining only one or two terms beyond p0 in the pressure series. In our case the second-order solution p2 does not contribute to the bearing load support, so that the load characteristics can be determined fairly accurately with the first-order solution pi alone. The First-order Solution. The first perturbation Eq. (5-45a) is made homogeneous by the substitution Pi = Pkoo) + Z(z)e(6) (5-46) where Z is a function of z alone, 0 is a function of 6 alone, and Pim is the corresponding term in the perturbation solution for an infinitely long bearing, or Pi(o°) = ^^2 (sin 6 “ A cos (5-47) When Eq. (5-46) is substituted into Eq. (5-45a), a homogeneous equa¬ tion in Z and 0 is obtained; it can be solved by separation of variables. The boundary conditions for our problem are: 1. For pressure continuity pi, and therefore 0, must be periodic in 8: Pi(z,8) = pi(z, 8 + 2tt) 2. From symmetry ph and therefore Z, must be an even function in z: = 0 at z = 0 dz
Hydrodynamic Gas Bearings 165 3. pi must vanish at the ends of the bearing: pi = 0 at z = ± By Eqs. (5-46) and (5-47) the third boundary condition can be stated in the form z(z= ± £) 9(9) = -plM = (A cos e - sin 0) (5-48) Equation (5-48) indicates not only that 0 is periodic in 0 but that it is made up of first-harmonic terms, sin 0 and cos 0. It can be verified by substitution that the first-harmonic solution to the homogeneous equa¬ tion in ZQ satisfying the first two boundary conditions is ZQ = Ci (cosh al cos j31 cos 0 — sinh az sin pi sin 0) + C2(cosh az cos pi sin 6 + sinh az sin pi cos 0) (5-49) , . Vl + A* + 1 where a2 = — 2 Vl + A2 - 1 2 and Ci and C2 are arbitrary constants which are determined by substitut¬ ing Z(z = L/D)Q(0) from Eq. (5-49) into Eq. (5-48), which expresses the condition that the pressure be ambient at the ends of the bearing. Once Ci and C2 are determined, ZQ is completely specified and the final form of pi is given by Anpa 1 + A2 |L Pi = sinh (aL/D) sin (PL/D) + A cosh (aL/D) cos (PL/D) . , . . iinh‘'(aL/D) + cos’ (flL/D) Smh ** * _ cosh (aL/D) cos (pL/D) — A sinh (aL/D) sin (PL/D) sinh2 (aL/D) + cos2 (pL/D) cosh az cos pi j sin 0 Tsinh (aL/D) sin (pL/D) + A cosh (aL/D) cos (pL/D) + [ sinh ^(aLjD) + cos2~(pL/D) ' cosh al cos pi — A _ cosh (aL/D) cos (pL/D) — A sinh (aL/D) sin (pL/D) sinh2 (aL/D) -f cos2 (pL/D)
166 Theory of Hydrodynamic Lubrication Second-order Solution. Although the second-order solution is too lengthy to be given here, it is easily shown that the form of p2 will be P2 = P2(oo>(0) + fi(z) + fz(z) sin 2d + fz(z) cos 26 (5-51) The functions /i, /2, and /3 are linear combinations of even functions of z formed from products of trigonometric and hyperbolic functions. The second-order solution p2(oo) of the infinitely long journal bearing is of the form P2(») = Go + d\ sin 26 -f a2 cos 26 (5-52) The important characteristic of the second-order solution p2 is that it contains no first harmonics sin 6 and cos 6. This fact, as shown below, means that p2 does not contribute to the load capacity. The load on the journal must be equal and opposite to the resultant pressure force acting on the journal. Components of the load are given by WK = — R- J^D dz p p cos 6 dd WT = R2 [L D dz f2rpsmed$ J -L/D Jo y It is clear from the d6 integrals that only those terms in p which con¬ tain first harmonics, sin 6 and cos 0, will produce any load. Of the first three terms in the pressure series only pi contains such terms. Thus WR ~ —R2 [L/.I*ndz [*V*Pi cos 6 d6 J —L/D JO WT = R2 fL/,D dz P'tpisin 0 d$ J —L/D JO By substituting Eq. (5-50) into the above integrals and integrating, we obtain WR A lrenpaRL 1 + A2 r (« - 0A) sin (2(3L/D) - («A + fl) sinh (2«L/Z))1 L (L/D) Vl + A2 [cosh (2aL/D) + cos (2/3L/Z))] J Wt ^ A rrmpaRL 1 + A2 _ (a ~ 0A) sinh (2aL/D) + (arA + |8) sin (2aL/D)~\ R L (L/D) \/l + A2 [cosh (2aL/D) + cos (2pL/D)] J Corresponding expressions for an infinite width bearing are obtained by letting L/D approach infinity in Eqs. (5-53). The resulting limits are WR oo A2 irenpaRL 1 + A2 W Too = A wenpaRL 1 + A2 (5-54a) (5-546)
Hydrodynamic Gas Bearings 167 Figure 5-14 shows dimensionless load and attitude angle against bearing number. This method of solution applies only to small eccentricity ratio, in general less than 0.3. The first-order solution implies that the deflection is directly proportional to the load. This is not true at high eccentricity ratio. To obtain good approximation to the pressure distribution, addi¬ tional terms should be used in the pressure series. Fig. 5-14. Isothermal bearing load and attitude angle vs. bearing number for first- order perturbation solution. It can be shown that the quasi-steady-state Reynolds equation with squeeze film velocity terms present can be solved by perturbation tech¬ niques in the same manner as the static-loading case discussed above. When such an analysis is performed employing first-order perturbation of € and e'} the results show that a correspondence relation exists so that WD = (1 - 2a')Ws where the subscripts D and S correspond to dynamic and static cases respectively. It can further be shown that correspondence relations exist between the derivatives of force with respect to displacement and velocity, which are where e = A icdt da o)dt dWR = 2 /dWr\ = (dWT\ de 1 — 2a' \ de )d \ de )s -I = _ 2 (dW«\ = o (*WR\ \ 1 — 2ol y de Jd \ de Js dW de (5-55a) (5-556)
168 Theory of Hydrodynamic Lubrication These derivatives are necessary in dynamic analysis, and they will be further discussed in Chap. 8. Figures 5-15 and 5-16 show the dimen¬ sionless radial and tangential stiffness vs. dynamic bearing number A* = (1 — 2a')A obtained from the first-order perturbation of e and In the absence of whirling, A* = A and the same figures are still appli¬ cable. By use of Eqs. (5-55a) and (5-55b) the other two derivatives with respect to velocity can be obtained. 1.0 A* *4-— 1/A* Fig. 5-15. Radial stiffness vs. dynamic bearing number for isothermal first-order perturbation solution. Fig. 5-16. Tangential stiffness vs. dy¬ namic bearing number for isothermal first-order perturbation solution. 5-13. Linearized ph Solution. An improved analytical solution which largely eliminates the defects of first-order perturbation is accomplished by linearizing the differential equation by setting the product ph of pressure and film thickness as the dependent variable. The resulting solution is called the “linearized ph” solution. Considering isothermal conditions and letting s = ph, we can rewrite Eq. (5-43) as , (dh , d2s\ 2 (d2h , d2h\ a Trds (dhds , dh ds\ ,|" / ds\2 , / dsVl = 8 (to te + T* ai) - h [ (te) + (ai) J (5-56) In the limit as U —> 0, the pressure approaches the constant ambient pressure (p —■► pa or s —> pah); while in the other limit as U —> oo, the product ph approaches a constant (s —> constant). In each of these two limiting cases, the terms on the right-hand side of Eq. (5-56) vanish identically. It seems a reasonable approximation then to neglect these
Hydrodynamic Gas Bearings 169 terms for all values of U. With this approximation, Eq. (5-56) reduces to hs(te2 + a?)-6>‘uVx = s (a? + 1?) The next step in the linearization is to approximate s where it appears as a coefficient by s « pah. The resulting equation is dh &s _ M/ ds (dVi <Mi\ dx2 +dz2 pah2 dx Pa\dx2+ dz2) ( ' Equation (5-57) is linear in s and can generally be solved by separation of variables once h(x,z) is specified. For example, if h = constant, as in a stepped-thrust pad, the equation is particularly simple because the coefficient 6nU/pJi2 becomes a constant and the right-hand side is zero. If h is a more complicated function of x and z, one additional simplify¬ ing approximation may be necessary, and that is to replace the coefficient 6nU/pJi2 by a constant 6nU/paC2. d2s , d2s 6/uU ds (d2h . d2h\ ( d^ + 6?-^dx = J,‘\d72+d?) (5'58) Eq. (5-58) still yields s = constant as U —> oo. As U —► 0 the coeffi¬ cient in question approaches zero anyway, and the solution will still be a good approximation in that limit. Substituting s = ph/paC = s/paC, dx = R dd, z = R{, U = Ray, and h — C( 1 + € cos 6) into Eq. (5-58), we obtain the dimensonless equation d2S d2S ds ap + dr~ATe~ ~£C0S® (o‘59) where A = (6/zwR2/paC2) = 3/2(/zo>D2/paC2). Boundary conditions are s = 1 + € cos 6 when f = ± (L/D) s is periodic in 6 s is an even function of f Equation (5-59) and its boundary conditions are very similar to the first-order perturbation-pressure equation and can be solved in the same way. The resulting solution is s = 1 + 2 (0d* sin 6 + 9$ cos e) (5-60)
170 Theory of Hydrodynamic Lubrication where gtf = 1 — A sinh af sin 0f + B cosh af cos 0f 02f = 1/A A cosh af cos /3f -f B sinh af sin 0f ^ _ A cosh a(L/D) cos P(L/D) + sinh a(L/D) sin &{L/D) sinh2 a(L/D) -f cos2 P(L/D) ^ _ A sinh a (L/D) sin ${L/D) — cosh a(L/D) cos 0(L/D) sinh2 a(L/D) + cos2 0(L/D) a2 = (\/r+T2 + l)/2 02 = (Vl + A2 - l)/2 The pressure is then j __i_ s*n ^ cos (5-61) J)aCs J)a Tii A V h 1 + € cos e L e or v = pW + Va ( C0S 9 1 + € COS 0 where p(1) is the first-order perturbation-pressure solution. The load components Wr and Wt, parallel and perpendicular to the line of maximum-minimum clearance, are given by wR = -** jpj* dt ffj p cos e de Wt = R2 dt PJ p sin 6 de which, upon substitution from Eq. (5-61) and integration, yield _ 2r«*p. 1 - VT=7> [+W (, A \ ^ - T" "VT^T* " J-*/* I1 " T+T** 2 1 - -s/l vr = ^<Si-77==T^ (5-62) wT- (1 - V. -.-) /“; (i^x, „,r) * = WTW | (1 - v~?) (5-63) where 1T*(1) and IfV0 are the first-order perturbation solutions. The load magnitude W = y/W R2 + Wt2 and the attitude angle 0 = tan-1 (Wt/Wr) obtained from Eqs. (5-62) and (5-63) are W = )T<» \ 1 ~ \/l - «2 Sin2 4>u> (5-64) € Vl “ €2 tan <f> = Vl c2 tan 0(I) (5-65)
Hydrodynamic Gas Bearings 171 where and <£(1) are the load magnitude and attitude angle obtained with the first-order perturbation. For completeness, W(l) and 4>a) are shown in Fig. 5-14 as functions of the bearing parameter A for various values of length-to-diameter ratio L/D. 5-14. Numerical Solution. By using numerical methods, results which are applicable over the complete range of eccentricity ratios are obtained. With proper grid size and convergence limit, these results are in closer agreement with experiments than are the previously discussed methods of analysis. A+i. / M j n (hbAj-'j ’(4wty-i, / A/ piJ ■■ p‘.i" 1 Fig. 5-17. Mesh representation. For steady loads, the Reynolds equation is by Eq. (1-11) d_ dx ( hz dp\ d ( h3 dp\ \ dx) dz \ n dz) dx (5-66) By using the dimensionless parameters x P P = — Pa and substituting them in Eq. (5-66), we obtain d (. h3 dp\ (D\ d / h3 dp\ d(ph) di\p^Tx) + {L) di[pJTz) = 6T-dx~ (5-67)
172 Theory of Hydrodynamic Lubrication Referring to Fig. 5-17, Eq. (5-67) can be written in finite difference form as follows: d_ dx d_ dz (^lB) Ky+i ~ fij _ /^*p\ Pi.i ~ Pi.j-1 /_ h3 dp\ = \ m /,, Ax \ jl Jj, j-u Ax \ pi dx) Ax (fcp\ Pi±hJ_-P}'J _ (^lI) BiJ ~ Pi-i.J (- !£ 12\ - V M A+k. j Az V P Ij-Yx, j Az V /Z dz) Az d (p^)t.y+H (p^)».i—w ai(p/i) Ai The pressure in the center of any grid is Pi.i = + tt; M(pfe)t-,y-H - (ph)i,j+u] , /^p\ p».i+l , / ^3p\ Pi, j-1 Ax \M Ay-m Ax2 \ M / j. j- m Ax2 _L + ^ l + _L \ A /i.i+14 \ jl /i.j-y. Ax* \L/ [\ jl /t+w,y Az2 + _L‘ V M A-^.y Az2 w r/** p.+i.j, A3p\ p»-i.>1 \L/ LVm Az2 \fl Az2 J +@L,A] Both the density and viscosity in Eq. (5-68) are functions of temperature and pressure. The pressure-density relationship is given below: V_ = /pY Pa \PaJ where n may take on values between unity (isothermal) and the ratio of specific heats (adiabatic). Under isothermal conditions p ~ pand/Z = 1. Since there is no film-thickness variation in the z direction, j = ht-Vi, j and Eq. (5-68) can be written in the following form: pij = Uo + a\pi+\tj + a2pi,j+i + dzpi-i.j + d\pi,j-\
Hydrodynamic Gas Bearings 173 € = 0.8 '0.6 1 _,i/ 6tt I S'WpnLRk .U^ p7 " LID--1/2 6 fiu (I? z~p~\'c A 0.4 0.2 '0.1 0 0.2 0.4 0.6 0.8 1.0 0.8 0.6 0.4 0.2 0 A 4- 1/A 1/A Fig. 5-18. Dimensionless force vs. bearing Fig. 5-19. Dimensionless force vs. bearing number. number. 1/A Fig. 5-20. Dimensionless force vs. bearing Fig. 5-21. Dimensionless force vs. bearing number. number. Fig. 5-22. Attitude angle vs. bearing Fig. 5-23. Attitude angle vs. bearing number. number.
174 Theory of Hydrodynamic Lubricalion 90 80 -e-70 £.0 ° 50 a> T3 £ 40 5 30 20 10 °0 0.2 0.6 1.0 0.6 0.2 0 A 4* t/A Fig. 5-24. Attitude angle vs. bearing number. 2.60 2.40 2.20 2.00 1.80 1.60 UH 1.00 0.80 0.60 0.40 Q20 °0 0.2 0.4 0.6 0.8 1.0 0.8 0.6 0.4 0.2 0 A *1* 1/A Fig. 5-26. Friction force vs. bearing number. Fig. 5-25. Attitude angle vs. bearing number. x Computer solution (Ref.12) ° Linearized ph solution (Ref.10) ^ Infinitely long solution (Ref.15) (end flow corrections(Ref.9) • First order perturbation solution (Ref.9) Fig. 5-27. Dimensionless force and atti¬ tude angle vs. bearing number, compari¬ son employing various methods of analysis. These equations now can be solved by relaxation, iteration, or matrix methods, and the result similar to the work in Chap. 4 yields a complete pressure profile in both x and z directions. Figures 5-18 to 5-26 give results obtained using a digital computer for the solution of isothermal gas bearings. Figure 5-27 compares results obtained by first-order perturbation, linearized ph, and iterative pro¬ cedure. The linearized ph gives an improvement over the first-order perturbation. Figures 5-28 to 5-31 give results of radial and tangential stiffness and damping vs. eccentricity ratio for several values of L/D and
Hydrodynamic Gas Bearings 175 Fig. 5-28. Radial stiffness vs. eccentricity ratio. Fig. 5-30. Radial force derivative with respect to velocity vs. eccentricity ratio. e' =“ dt/ta dt Fig. 5-29. Tangential stiffness vs. eccen¬ tricity ratio. '■o I 30|- D 20 4 = 2 ■S^l.5* A A = 0.5 o A= 1 - - = 1.5- k- f - 0 0.2 0.4 0.6 0.8 1.0 € Fig. 5-31. Tangential force derivative with respect to velocity vs. eccentricity ratio, e' = dt/u dt compressibility number A. These analysis and are discussed in Chap. functions are necessary for dynamic 8. SOURCES 1. Burgdorfer, A.: The Influence of the Molecular Mean Free Path on the Per¬ formance of Hydrodynamic Gas Lubricated Bearings, Trans. ASME Ser. D, vol. 81, March, 1959. 2. Constantinescu, V. N.: Consideratii Asupra Calculului Lagerelor de Alungire Infmita Lubrificate cu Gaze, Compuse din Suprafete Plane (Methods for Calculating Characteristics of Plane Surfaces and Lubricated by Gas), Studii Se Cercetari de Mecanica Aplicata, vol. 7, no. 3, 1956.
176 Theory of Hydrodynamic Lubrication 3. Harrison, W. J.: The Hydrodynamical Theory of Lubrication with Special Reference to Air as a Lubricant, Trans. Cambridge Phil. Soc., vol. 22, no. Ill, p. 39, 1913. 4. Gross, W. A.: Compressible Lubrication of Infinitely Long Slider and Journal Bearings, IBM Notes, June, 1958. 5. Brody, S.: Solution of Reynolds Equation for a Plain Slider Bearing of Finite Width with an Isothermal Gas Flow, ASLE Paper 60AM5A-4. 6. Osterle, J. F., and W. J. Hughes: High Speed Effects in Pneumodynamic Journal Bearing Lubrication, Applied Scientific Reseach, Section A, vol. 7, 1958. 7. Sheinberg, S. I.: Gas Lubricated Journal Bearings, Izvest. Akad. Nauk, S.S.S.R., vol. 3, pp. 107-204, 1953. 8. Katto, Y., and N. Soda: Proc. Second Japan. Natl. Congr. Appl. Mech., 1952. 9. Ausman, J. S.: The Finite Gas-lubricated Journal Bearing, Conf. on Lubrication and Wear, Paper 22, London, 1957. 10. Ausman, J. S.: An Improved Analytical Solution for Self-Acting Gas-Lubri¬ cated Journal Bearings of Finite Length, ASME Paper 60-LUB-9. 11. Sternlicht, B., and R. C. Elwell: Theoretical and Experimental Analysis of Hydrodynamic Gas Lubricated Journal Bearings, Trans. ASME, vol. 80, June, 1958. 12. Sternlicht, B.: Gas Lubricated Cylindrical Journal Bearings of Finite Length, Part I, Static Loading, Aug. 15, 1960, ONR Report, Contract No. NONR 2844(00). 13. Sternlicht, B.: Gas Lubricated Cylindrical Journal Bearings of Finite Length, Part II, Dynamic Loading, Sept. 14, 1960, ONR Report, Contract No. NONR 2844(00). 14. Raimondi, A. A.: A Numerical Solution for the Gas-lubricated Full Journal Bearing of Finite Length, ASLE Paper 60LC-14. 15. Elrod, H. G., and A. Burgdorfer: Refinement of the Theory of Gas Lubricated Journal Bearings of Infinite Length, Proc. First Intern. Symposium on Gas Lubricated Journal Bearings, Washington, D.C., October, 1959.
CHAPTER 6 HYDROSTATIC BEARINGS The pressurized or hydrostatic bearing, in contrast to the hydrody¬ namic bearing, demands external fluid pressure to support a load. This system of lubrication offers some distinct advantages not found in hydro- dynamic bearings. Among the most important of these characteristics are extreme rigidity, a load capacity independent of velocity, and very small frictional drag. Fig. 6-1. Pressurized bearing operation. In the hydrostatic bearing, the fluid film that separates the journal and bearing is maintained by a source of pressurized fluid external to the bear¬ ing. The pressures in this fluid are regulated, or compensated, by flow restrictions both in and external to the bearing. Figure 6-1 illustrates the principle of pressurized bearing operation. Fluid at pressure p, enters the bearing clearance through external restrictors R9X, which can be either orifices or capillaries, and exhausts through the bearing clearance restrictions Ri and Ri. When the journal is centered in the bearing, the pressures pi and pi in the upper and lower portion of the bearing, respec- 177
178 Theory of Hydrodynamic Lubrication tively, are equal and no net force is exerted on the journal. However, when the shaft is deflected to increase the clearance hi, the restrictions Ri and R2 decrease and increase, respectively, thus causing pi to decrease and p% to increase. This action may be visualized with the aid of the electrical bridge shown in Fig. 6-1. A net force, proportional to the shaded area in the pressure distribution diagram, acts to oppose shaft deflection. Thus, the bearing load-carrying capacity is a function of the rate of change of pressure with clearance. Although the frictional losses in the hydrostatic bearing are quite low, the total power required for the operation of hydrostatic systems may be very high, because considerable power is required for pressurizing the fluid. Thus efficient design requires optimization of the supply flow relative to the load-carrying capacity. Both laminar and turbulent flows are considered, and, as will be seen later, from the standpoint of power requirements it is preferable to maintain turbulent flow in the restrictors. Consideration is given to both compressible and incompressible lubricants, and conditions for stable operation with compressible fluids are presented. The assump¬ tions made are that pressure drop from the recess to the bearing periphery is linear and that, for small deviations from equilibrium, this type of pressure distribution is maintained. In the presence of relative motion between the two surfaces a combina¬ tion of hydrodynamic-hydrostatic lubrication results. The effect is to raise slightly the load-carrying capacity. PLAIN JOURNAL BEARINGS 6-1. Incompressible Lubrication. Laminar Feeding. For the non¬ rotating bearing the pressure distribution may be expressed by setting U = 0 in Eq. (1-12) as (6-la) or d/p d/p 3 dhdp dz2 dx2 h dx dx 4. ~ _ rr = 0 (6-1 b) By writing x = R6 and h = C(1 — € cos 6), we have 3 dh _ 3e sin (x/R) h dx C[1 — e cos (x/R)] The boundary conditions for this problem are p = Pa when z = ± — p = pi(x) when z — 0
Hydrostatic Bearings 179 where subscript i represents inlet to the bearing clearance and where q is the flow per unit length. If the fluid enters the bearing through a laminar feeding tube, the pressure drop in the tube may be written as P. - P> = (6-2a) where I is the length and a the radius of the feeding tube. Let N be the number of feeding tubes; then the number of feeding tubes per unit length in circumferential direction is n = N/2irR and the above equation becomes = wip,~Pi) Thus from continuity Ta*n ( \ h* (dP\ ta on 9“ = W(p‘-pi) = "sfej, (6-26) By substituting the film-thickness expression in Eqs. (6-2), we have 3 ira4n(pt — pj) 4C3/[1 - c cos {x/R)Y (6-3) This equation represents the inlet boundary conditions into the bearing. Since dp/dx —► 0 as c —> 0, for small c, dp/dx is of the same order of magnitude as e. This means that (3/h) (dh/dx) (dp/dx) is of the order c2 for c 1.0, which allows Eq. (6-16) to be written as 3+S- *«'•» By separating variables in Eq. (6-4), we have p = X(x)Z(z) which yields two ordinary differential equations H + K.X.° g-K.Z-0 where if is a constant. This gives p(x,z) as (6-4)
180 Theory of Hydrodynamic Lubrication sin (jx/R) having been disposed of on the basis that pressure distribution must show axial symmetry. Since e cos (x/R) 1.0, we can use in Eq. (6-3) the binomial expansion m 1—Tip vf3 = 1 4" 3c cos "d 4“ 6 (c cos -5j 4~ • • • [1 — «cos (x/R))z R \ RJ Thus the boundary condition given in Eq. (6-3) is to a very good approxi¬ mation (!j). = -Ai ^1 + 3e cos-|) (p. - Pi) (6-6) where only the first-order, or linear, terms in c have been retained and _ 3to4n L " ~icn Applying boundary conditions to the plus z half of the bearing gives C.--S* B, = — AjeiLIR so that Eq. (6-5) becomes 00 p(x,z) = Pa + Cl (z - 7j) + A> COS'! {e’‘IK ~ e,<L"‘)/fi) (6‘7) 3 = 1 From which Pi = pa - Ci ^ 4- ^ Ay cosJ~ (1 - e’LIR) 3 = 1 (6-80) and (S). = + 2 ^ (i) ^ j-1 From Eqs. (6-85), (6-6), and (6-8a) we obtain 00 c\ + J A, cos! (!) (1 + e*!«) = -A,. (1 + 3« cos |) 3 = 1 00 [p. - [p. - 2 A> cos! (1 “ e'L/ft)]) (6-0) 3=1 from which r - - A^(P« ~ ?«) 14“ AlL/'2, 4 _ 3A lR(p, — Pa)e /1 1 a / 7oui i '/./«' i i (l + A,.L/2)[1 + eL'R + ALR{eLIR - 1)]
Hydrostatic Bearings 181 and, neglecting higher-order terms of e, A2, A3, . . . , Aj = 0. Upon substitution of these constants into Eq. (6-7), the pressure distribution becomes p(x,z) pa + (2 *) 3Atft(p. - p.Mcos (x/ft)(e"« - e<L-'»'*)] (1 + AtL/2)[l + eL'R + ALR{eL'R - 1)] for e 1.0 The load-carrying capacity is (6-10) W = 2 JqL/2 J2r p cos OR dd dz By using Eq. (6-10) and integrating, we obtain w = forAlR\v* ~ p0)[cosh (L/2R) - l]c . . (1 + AlL/2)[cosh (L/2R) + ALR sinh (L/2R)] p _ W _ ZirALR*(p. - Pa)[cosh (L/2R) - l]e 0r LD L( 1 + AtL/2)[cosh (L/2ft) + At* sinh (L/2ft)j The power required to pump the fluid through the feeding line into the bearing is h = (p.~ p„) j\„Rde = -(P. - po) /o2'£(g)(ft^ From Eq. (6-86) in which the constant C\ is substituted we have for the pressure gradient (£), - - tTxim + X ^ C“^/R)l1 + «"■> (6-12) i-i The expression for film thickness may be expanded to h3 = c3(i - € cos ey = c3[i - 36 cos e + • • •] The product of the last two equations integrated with respect to 0 over the interval 0 to 2x and with terms of order «2 and higher neglected yields _ rRALC3(p. - pa)3 Hl ~ 3m(1 + AlL/2) ' (6_13) in which it should be noted that H is independent of t for small values of (. Eliminating p, — p0 between Eqs. (6-11) and (6-13) gives 4 C2P2(L\ L 27*pt3\2R) 2 (1 + At L/2) [cosh (L/2 ft) + At ft sinh (L/2 ft)]2 Atft[cosh (L/2ft) - l]2 (6-14)
182 Theory of Hydrodynamic Lubrication Minimizing H with respect to Al reveals A _ J_ JTi I *L/R 1* 2L tanh (L/2 fl) J Equation (6-15a) can be rewritten as f1 , 8L/D f _ j (Na*\ _ L tanh (L/D)\ \ 1C* /0p» 1.5L/L By letting the minimum pumping power for an arbitrary L/D ratio be defined as the reduced power, we have H - Hl «red — (6-15a) (6-156) Cip* 108 TM€2 = kU + (l+ 8V*> Y\ D I V tanh L/D) j (4(L/D) cosh (L/D) + [(l + ~ l] sinh (L/D))* [(' + - •] (t/B) - 1J= (6-16) By taking the derivative of HT9d with respect to the L/D ratio, we obtain a minimum value Hrtd of 721 at L/D = 1.1. At this point Eq. (6-16) yields ftp* ^ = 2.15^- (6-17) Solving Eq. (6-11) for L/D = 1.1 yields (p. - P«)op. = 2.43^ (6-18) The shear stress by Eq. (l-15a) is /du\ ( U 1 dp ,\ Since for small e, dp/dx is of the same order of magnitude as €, the fric¬ tional forces are given by , , 2tthRLU lft, Txdxdz = ~— (6-19) CL/2 f: = 2/o Jo Turbulent Feeding. The Reynolds number for a capillary can be defined as Re = 2 apw = 2 Qp jl 7T a/X
Hydrostatic Bearings 183 where a is the radius of the feeder and w the average velocity. Since for laminar flow Q = C»Ai,(p. - p.) v 6/*n(l + AtL/2) the Reynolds number becomes (6-m> For the minimum power loss, the above reduces to C*pP Re = 0.56 eNap2 If Re < 2,320, laminar flow is maintained; if Re > 3,000, turbulent flow is experienced. Turbulent flow can be achieved in one of two ways: for turbulent pipe flow, at Reynolds number greater than 3,000 the pres¬ sure drop is given by fl 2 Va-Vi = -pW* where/is the pipe friction factor. The volumetric flow rate is thus given by Q = (V« ” Pi) For series orifice flow the pressure drop across an orifice is given by pw2 Pi - Pi± i - 2C? where C* is the discharge coefficient which is generally in the neighborhood of 0.63. For m orifices in series P. - Pi = Tjjjj and Q = wa*Cd y]^(P> ~ P.) It has been shown that so that, neglecting terms containing higher orders of e, we obtain (1), = ~At 0 +3e cos i) v?‘ ~ p* where subscript T stands for turbulence under incompressible condition.
184 Theory of Hydrodynamic Lubrication Thus for turbulent flow in a pipe 12 ira2pn Ya ' v/£ and in an orifice Let By using the previous boundary conditions, we obtain for the integration constants /nr __ Pa) (L/2)(1 + Xr) 3cXr (p$ — pa) 2(1 4- \t)[(L/D) cosh (L/D) 4- 3^Xr sinh (L/D)]eLlD and as before Ai} Ai} . . . , Aj = 0. By substituting these constants into Eq. (6-7), we get 3<Xr(e‘!R - - p.) cos (x/R) . . 2eL'®(l + \t)[{L/D) cosh (L/D) + (Xr/2) sinh (L/D)J K ’ Now, by comparing this result with the result previously obtained for laminar feeding where Xl = AlL/2, we have i Mp.-p.) P P° + 1 + XL M) 3e\L(e*lR — e{L~t),R)(pt — pa) cos (a/fl) 2ei'"(l + Xl)[(L/D) cosh (L/D) + Xl sinh (L/D)] (6'23) The last expression differs only in form by the coefficient of sinh (L/D), which is 0.5X7- for turbulent flow and Xl for laminar flow. The dimen¬ sionless quantities Xt- and \L have, however, different values. For laminar flow 37r a*Ln f . Xl = (6-24) For turbulent capillary flow <«*>
Hydrostatic Bearings 185 For series orifice flow ^ mC«P(p. The load-carrying capacity is obtained by integrating Eq. (6-22): p 3irXr[cosh (L/D) - l](p, - p„)i . . 2L/D(l + \t)[(L/D) cosh (L/D) + y2\T sinh (L/D)) y > By defining p = — V. = & Pa€ Pa we get for turbulent feeding 5 = 3?rXr[cosh (L/D) - 1 \(p, - 1) . (2L/D)(1 + \t)[(L/D) cosh (L/D) + sinh (L/D)} ° and for laminar feeding p = 3TXL[cosh (L/D) - l](p, - 1) __ (2L/D)(1 + \l)[(L/D) cosh (L/D) + \L sinh (L/D)j V ; The power necessary to pump the fluid through the bearing is H = (p. - Pa) f/'q„Rde where for turbulent feeding the flow per unit circumferential length is q- = fr (xttt) (1 +3t cos ~ JJ ttC3D ( Xt \ ( ^ HT = ^L\yyXr) {P‘-V°y so that Let the dimensionless power be defined by f{T = _ Tr\TD(p, — l)2 C*p o2 3L(1 + \T) then elimination of p8 — 1 between Eq. (6-28) and the last expression yields Q _ 4 P*L(\T + 1 ){(L/D) cosh (L/D) + y2\T sinh (L/D)Y T 27ir CXr[cosh (L/D) - I]2. ^ J As a comparison the dimensionless power for laminar feeding is 0 _ 4 P2L(\i + 1 )[(L/D) cosh (L/D) + \L sinh (L/Z))]2 L 27ir DXl[cosh (L/D) - I]2 ( '5U)
186 Theory of Hydrodynamic Lubrication Minimizing S with respect to X, that is, setting dH/d\ = 0, gives the optimum values for X: f'T opt = i (Ti + m ?_i| 4 j [ £> tanh (L/ D) J ‘j Xl"pt = ijl1 + DtoHhWZT)] " *} Insertion of these values in Eqs. (6-29) and (6-30) yields l([1 + P t>nh «,/P)] +3|(sC“hS +n[i+i> jt/p)r-ii“"b^p)' |[‘+ Dtafwcif - 1)(cosh B -') (6-31) B |[' + P imMI/Pl]* + 3I(b I°*h^ J ([> + PtenhV/Dir-1!”111*^) IP ^ p2 4 Hl ~ 27iP “ If1 + -mJh(L7D)T ~1 |(cosh i-1) (6-32) Evaluation of above equations gives the following optimum parameters 104 #rmi„ = 1.45P2 (IL= 0=^ ^ = 1.10 ffL^ = 2.15P2 (6-33) (XT).!. = 0.934 (P*)rop, = 1 + 1.725P (Xz.),,p» = 0.615 = 1 + 2.43P From this information it can be shown that for turbulent capillary flow /oW\ =0Q2p V/^w.P. and for turbulent orifice flow /.WWN ,0Mp \ mC«ppa /opt while for laminar flow / aAN\
Hydrostatic Bearings 187 Rotational Considerations. The velocity components in a fluid film were derived in Chap. 1 as 17 o -1) - s is ^ ^ If we find the average value of the velocity through the fluid film and combine the two equations into one vector v, we get 1 tt h2 _ T = 2 12m and for the continuity equation V • (hv) = 0 Under isothermal conditions the above equations yield U(Vh) - ~ (Vh) . (Vp) - Jji Ap = 0 If the z axis is taken as a line generating the cylindrical journal surface, taking into account that dh/dz = 0, the last equation gives dh ^Wdhdp + ^P\ /6_34) dx 2m dx dx + 6m \dz* + dx7 { ’ We will limit ourselves to the case where the eccentricity ratio is small and second or higher powers of e can be disregarded. Thus Eq. (6-34) becomes Cell . x — Xo Ch . X — Xodp C3 (d2p . d2p\ 2rn^- = ssm"irari2i^ + ¥‘) (6-35) where xo = R<f>. However, as explained previously, dp/dx is of the order of magnitude of «, and we have d2p , d2p _ 6fiU x - xo ,n dx2 + dz2 'RC2 * R ( ^ One boundary condition is z — L/2, p — pa; the other boundary con¬ dition follows from the consideration that the quantity of lubricant that flows through the capillaries is the same as that leaving the bearing. This latter condition leads to 37r a4n p, — pi 4/C3 [1 — e cos (x — Xo)/R\z
188 Theory of Hydrodynamic Lubrication To solve for p in Eq. (6-36), we assume the solution is of the form p = X(x)Z(e) + A sin By substituting the above expression into Eq. (6-36), we get Here K2 must be a positive constant because X must be a periodic func¬ tion with a fundamental period of 2tR. The individual values of K are therefore K = j/R (j = 1, 2, . . . , n). Lastly, since Ciz -f C2 rep¬ resents a specific integral of the homogeneous differential equation, the result will be of the form V = P. + c* + c2 + £ (Au*<* + At,e~’,IR) cos j i-i + £ + B2ie-“'*) sin j ^ tR sin (6-37) >- 1 If we substitute the two boundary conditions and disregard terms containing the second or higher power of e, we get An = A2j- = 0 j = 2, 3, . . . , n B\j = Bn = 0 j = 2, 3, . . . , n n ^ _ MP* Pa) //» oq\ 2 ~ "T+r (6_38) r _ A(p, - Pa) 2 1 + X 3A(p, — pa)e-LIDe Aw — 2[(L/D) cosh {L/D) + <r\ sinh {L/D)]{ 1 + X) 4 = 3X(p, — pa)eLlDe *21 2[(L/Z>) cosh {L/D) + a\ sinh {L/D)]{ 1 + X) 6pUR[L/D + a\{ 1 - e~L,D)]e Bn = C22[{L/D) cosh {L/D) + *A sinh {L/D)] 6nUR[L/D + <r\{eLiD - l)]e “ C22[(L/Z>) cosh {L/D) + (tX sinh {L/D)] X above is defined by Eq. (6-24) for laminar flow and Eq. (6-25o) for turbulent flow, while <j is unity for laminar flow and 0.5 for turbulent flow. By substituting these equations in Eq. (6-37), the pressure distribution can be evaluated.
Hydrostatic Bearings 189 To determine the shearing stress distribution across the journal surface, we start with the fundamental equation (l-15a): T = _ * *2 _ £ t; * 2 dx h By substituting constants from Eqs. (6-38), we get r* = — ^ ^1 + e cos —+ A2ie~*IR) sin X—^X° - (Bne-'s + Bne-i«) cos + ^(R cos L^°] j (6-39) In order to determine the position of the journal with respect to the bearing, we must introduce the equilibrium requirements for the journal. The only inertia force acting is that due to gravity. Therefore the follow¬ ing equilibrium conditions are applicable: W = 2R j^/2 dz f2* p cos Odd -+• 2 R j^/2 dz j2* ts sin 0 dB (6-40a) r l/2 r 2w r l/2 r2w 0 = — / dz / p cos 0 dd + / dz / rx cos 0 d0 (6-405) These two equations yield two relationships for the two unknowns e and <f> that will establish the position for the journal. Neglecting all terms of order of C/R and integrating Eq. (6-40a) yields P = eE(cos <f> + 7 U sin <£) (6-41) whprp F - 3T\[cosh (L/D) - l](p. - pa) 2(X + 1 )(L/D)[(L/D) cosh (L/D) + *A sinh (L/D)\ 2 ^ C2(pt — Pa) (1 + X){cosh (L/D)[(L/D)2 - 2<rA] + (L/D) sinh (L/D)(a\ - 1) + 2<rA| X[cosh (L/D) — 1] If we substitute in this the optimum values for laminar and turbulent inlet flow, we get E = paP Rn C^aP Ryi nw, = 2.04 (6-42) Eq. (6-406) yields yT t = 1.08 r°p‘ C'pJ> cos <t> = , *■ — sin $ = 2^.—= (6-43) Vl + Vl + y2U* The position of journal center can now be obtained from Eqs. (6-41) and (6-43).
190 Theory of Hydrodynamic Lubrication For a stationary journal 0 = 0, U = 0, so that the center of the journal is exactly below the center of the bearing. As U increases, 0 also increases. For the limiting case of U = oo, 0 = 90°. To find the shaft locus for different values of U, we will set up a coordinate system { and rj, such that £ = e sin 0 r\ — e cos 0 If we eliminate the parameter U from Eqs. (6-41) and (6-43) by using the new coordinates, we get *■4-3)'-(0)* This is the equation of a circle with its center on the rj axis and radius = (3) This passes through the center 0 of the bearing, but only reaches it for U = oo. If e0 is the amount by which the center of the journal is lowered for U = 0, we must have e0 = PC/E, or e0 _ 2 P (1 + \)[(L/D) cosh (L/D) + <rX sinh (L/D)) ( . . C ~ 3t p, - pa X[cosh (L/D) - 1] The friction torque can be obtained from the shearing stress distribution equation (6-39). We get M* = 2 (2* [Ln R2tx dedz = - ~(6-46) 6-2. Compressible Lubrication. Laminar Feeding. The Reynolds equation for compressible, laminar flow is given by Eq. (1-66). By using the perfect-gas relation for isothermal flow, p/pg = const, and since V = Ui = U2 = 0, we have h3lz(PJz) + Tx(ph3f/) = 0 (6-47) This equation when expanded becomes 1 r/apV -I- (^1Y1 -i- (^P 4- ( sin (x/R) 1 dp = . V L \^x) \^z) J \dx* l — c COS (x//2)J J dx As we did previously, we set tdp/dx ~ 0 and we have j[(!?MS)’MB+20=» Ml * Refers to frictional torque.
Hydrostatic Bearings 191 We may now define a new dependent variable p such that p = Vp which allows Eq. (6-48) to be written as 3+S=° ^ which is the same differential equation as for the incompressible case. If we put pa = pa2j the first boundary condition of our problem reads: for z = L/2 (from capillary exit to gap) p — pa. This is again the same condition as for incompressible flow, with pa replaced by p. Therefore, Eq. (6-23) is a solution to Eq. (6-49), except that p is replaced by p. Thus * = * + - ft) V Pa+ 1 + X, (-?) 3e\i(e'lR - e(L~‘)lR)(p, - pa) cos (x/R) . . 2eLiD(l + \i)[(L/D) cosh (L/D) + X, sinh (L/D)} ^~OV} By substitution of the relation between p and p, Eq. (6-50) may be written (neglecting higher-order terms in c) as „ _ „ T1 4- hp.2 - HP.2 - l)2g/Ll* ° L 1 + Xi J 3X/€(p#2 — \)(e*,R — eiL~t)IR) cos 6 \ | 4e*'*[l + p.2h - (p.2 - l)X,2z/L] I [ [(L/D) cosh (L/D) + \i sinh (L/D)]J (6-51) If we carry out the integration, the load capacity is p _ 3yXf6p0(p<2 — 1) 4 L Vf+ h eL‘D[(L/D) cosh (L/D) + X, sinh (L/D)} 1/2 (eVr-M)IB _ e*//2) dz 0 [1 + \ip.2 - HP.2 - 1)2*/Lp* s: (6-52) This equation may be integrated by using the transform _ \(L/D)[\ + \ip2 - HP.2 - D2*/L]|» I HP.2-i) J This integral can be resolved into the Gaussian error integrals 4>(z) = —= f e~1' dt and xf/(z) — f e+fl dt VnJo Jo
192 Theory of Hydrodynamic Lubrication both of which are tabulated. Thus from Eq. (6-52) we get e-r'MVY* + L/D) - *(F)] b_3t( - (Vx/2)eH*>(V(Y' + L/D) - 4>(F)]I 4Y | (L/D) cosh (L/D) + X, sinh (L/D) j ( l The isothermal pumping power required is H = gp q,R d$ PQ Pa JO Neglecting higher-order terms in «, the laminar pumping power is given by U _ *'Xip.sC',(p,8 - 1) ln p. ,a K<n Hl L/D^(i + X|) (6‘54) In dimensionless form this may be written as a _ r X,D(p.2 - 1) ln p. , , Hl ~ 6 —L(TTT,) (6 ) In terms of Y the above equation reads Hi = I^ln[l + + Xl) ] (6-556) If for brevity we put Eq. (6-53) into the form F (r- b) - m I*"” [* - 'H - ^ er’ [* (^jF2 + 75) - *(F)]J (6-56) then Eq. (6-53) can be written as p = F(Y,L/D) cosh (L/D) + \,D/L sinh (L/D) y 1 By eliminating X/ between Eq. (6-556) and the last expression, we obtain a _ ,n r J p sinh (L/D) _L_ "I ' 12 F2 [ F2 F(Y, L/D) — P cosh (L/D) L>Y2 J Figure 6-2 shows a plot of F(Y, L/D) against Y for various values of L/D. (Xdopt can be obtained also from Eq. (6-57), and it takes the following form: ©. AN - I u \ (L/D)opt] - P cosh (L/D)opX ,fi (M^-ln)^ P sinh (L/D), ' ( )
Hydrostatic Bearings 193 Finally, (p,)opt can also be obtained from Eq. (6-52), and it yields ».). p, = [ 1 + (s)opl VJ ] (6-60) By using the above equation, Fig. 6-3 was plotted; it compares the per¬ formance characteristics of laminar-fed bearings using compressible and incompressible fluids. 0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 r Fia. 6-2. Performance of laminar-fed Fig. 6-3. Performance of compressible compressible journal bearings. and incompressible laminar-fed journal bearings. Turbulent Feeding. The mass flow rate M may be obtained from MN = J*’ pq.R d9 If we neglect higher-order terms in e, we get MN = tX'P<.*<?3-P(P.2 - 1) 12?<R7VL(1 + X,) The Reynolds number may be expressed as He — ^aPw _ 2 A/ M Tray, Re = J&DPW --1)- 6iV(l + \,)La^g(SlT For Re > 3,000 we have turbulent flow. However, the exact treatment of turbulent flow in capillaries leads to difficulties, because inertia forces
194 Theory of Hydrodynamic Lubrication are necessarily involved. We can arrive at simple relationships if we take the choking tube as a basis of the calculations. The choking tube is one with a relatively large internal diameter in which there are a number of orifices in series whose diameters are 2a. If there is a sufficiently large number of these orifices, the density variation and temperature variation for flow through each individual orifice need not be considered. Since the kinetic energy is dissipated between the orifices, the pressure drop across any single orifice is Aft From continuity the mass rate flow is M — ra2p/Wj and the equation of state for isothermal conditions is Vi/PiQ = (RT = K Combining the above relations yields „ A_ _ KM2g Pi Vi 2ir2CVo4 By summation over all the orifices, we obtain mKM2g t ft A Vi = If there is a sufficiently large number of orifices, the sum on the left- hand side is approximately equal to a known integral and we obtain ‘ V dv = - (v 2 - v 2) = mKM- g- pdp 2 (P. p.) 2w*C/at By defining p = \/p, the mass flow rate per feed tube is If n is the number of feeding tubes per unit circumference, continuity of flow leads to the equation -ip‘(s).=^a2(^r (d/\ = \dz )i (1 — € cos — Vi) ^ cos 6y - Pa , . 12\nrCdd2n (Kg\54 2\t lpa where A, ^ ^T+ M Subscript t stands for turbulence for compressible fluids.
Hydrostatic Bearings 195 By using the boundary conditions p = p« at z = L/2, the solution for the pressure distribution has the same form as in the laminar-feeding case, namely, Eq. (6-51) „ _ „ r 1 + M.2 ~ HP.2 ~ l)2g/L"|* p p° L i + J ( 3X«<(p,2 - l)(e'/s - e{L~‘)IR) cos 6 ) j 4eL/D[l + p.2X, - (p.2 - l)X,2z/L] } (6-61) { ((L/D) cosh (L/D) + (X,/2) sinh (L/D)]J By defining the mean bearing pressure in a manner analogous to the laminar feeding case and by integrating Eq. (6-61), we have le-y'lHVWTL/D) - *(F)J p = *1 j - ^ + L/D) - *(y)JJ (6.62) ( (L/D) cosh (L/D) + sinh (L/D) In terms of parameter F the above equation becomes p _ nr, W) (6.63) cosh (L/D) + (X,/2L)D sinh (L/D) V ’ where F(Y, L/D) = fg je- [* (^F2 + £) - *(F)] - ^>k,KVf2+5)-$(f)]| The dimensionless pumping power is analogous to Eq. (6-556) [1 + (L/Zx,K2+Xl-] (6'64) T I2F2 J By eliminating X* from Eq. (6-64) and (6-63), we obtain an expression analogous to Eq. (6-58): fj t i [~ I P sinh (L/D) L 11 ‘ 12F2 ln [2F2F{Y, L/D) - P cosh (L/D) + DF2 + 1J (6‘65) For a given value of the quantity P we can find a minimum for with respect to the variables Y and L/D. In this way the values of (L/D)opi and Fopt are obtained as functions of P; the related values of Hmi1x are also obtained. (X<)opt then follows from Eq. (6-62): _o/-\ F(Fopt, (L/D)opt) - P cosh (L/ D)opt (Q P sinh (L/D)0
196 Theory of Hydrodynamic Lubrication 12 10 (i\ r u "1—7" - n) Tf i L ii: tzi V - £ V -4- rrxl \t _ — i o!,.. 7 A _L h n~ —h- s s — >£ / f- L - Ht- r** W'\ -S- T -l ^up' T- i fimum turbulent incompressible :ompressible t = < 1.2 L D Optimum power 'Turbulent T - incompressible^ t - compressible Lominor Fig. 6-4. Performance of compressible Fig. 6-5. Optimum pumping power re- and incompressible turbulent-fed bear- quirements in journal bearings. ings. Figure 6-4 compares the bearing performance characteristics of the com¬ pressible and incompressible turbulent-fed bearing; Fig. 6-5 compares dimensionless pumping power for the various cases studied. The fric¬ tion force is similar for all four cases and is expressed by Eq. (6-19). STATIC AND DYNAMIC CHARACTERISTICS OF GAS JOURNAL BEARINGS The dynamic analysis of journal bearings can be considerably simplified if we replace each symmetrical portion of the journal bearing by an equivalent thrust segment and neglect A.M.* crossflow as shown in Fig. 6-6. A m further simplification is that the pres- -^C+e sure gradients are linear. With these simplifications the change in force on the full bearing AF is related to the change in force in the model AF0 by the equation -* Lp -*• -I - \ AF = ki AF0 (6-68) Fig. 6-6. Simplified model of a pres¬ surized journal bearing. where the constant ki normally is equal to one-half the number of lubri¬ cant inlets in the bearing. If the
Hydrostatic Bearings 197 actual pressure distribution in the pad is replaced by a constant circum¬ ferential pressure and a linearly decreasing axial pressure, the change in pad force AF0 as a function of change in pad pressure Ap0 is found to be Af o = 6(L + Lp> Ap„ (6-69) and the corresponding shaft force is Af = k, b(L Ap0 = A. Apa (6-70) where Ae is an effective area over which Ap0 is assumed to act. The flow into and out of the pad is controlled by the combined valving action of the inlet and the clearance-space restrictions. If we assume that the eccentricity is small, the incremental-flow relations for the equivalent pad can be written in the form AM, M, AM, ria _ f c dMjA e /Poi dM{1\ Ap0 \MidCjC \MidpoJpoi U} out / ^ £ | /Poi ^^fout\ Apo /p " \Mi dC JC* \M> dpQ J poi M, where p0i and Mi are respectively pocket pressure and the mass flow through the pad when e = 0. For simplicity Eqs. (6-71) and (6-72) are written in the form A Min , e Apo -M7 = hlc~a^ AM.,, u e , _ Ap0 -M~ = h'C + a^ The quantities «i, a2, &i, and 62 are effective valve parameters for the bearing. The net flow into the pad is AMia —- AM0Ut \ £ / i ^ Apo (p — = _ 52) _ _ (ai + a2) — (6-73) The conservation of mass requires that the net mass flow rate into the pad equal the rate of change of mass inside the pad M in — Mout — dM,md dt d dt d dt PoLpbpA + PobLp(C + e) + f6(L- LP)(C + e)j po [V, + b(L + hp) (C + «) j J (6-74)
198 Theory of Hydrodynamic Lubrication where po is the pocket density and Vp is the pocket volume. By lineariza¬ tion of Eq. (6-74) for small changes from the steady-state conditions e = 0, po = poi and dpQ/dt = 0 yields A M{n — A Mout = p«hl*Lr)% + \ v> + HLtLp)c]ii (6_75) Since the changes occurring in fluid properties inside the pad are small and are likely to be rapid in comparison with thermal lags, we can assume an isentropic relation between p and p. Thus ^-° = k ^ (6-76) Poi Poi Combining Eqs. (6-70), (6-73), (6-75), and (6-76) gives us = ” i jl [-(L +2Ly)CM‘] + (62 - 60 j (6-77) Equation (6-77) can be rewritten in the form (6-78) where kd is the dynamic spring stiffness: A F kd = - — (6-79a) e k, is the static spring stiffness: \bz — bi , _ Atp,/poi\l ‘ C \p.) i ' 0L\ -|- 0,2 n is the lead time constant: Poib(L + Lp)CMi (6-79 b) (6-80 b) '■ - 2«, -Ij ■ ' t2 is the lag time constant: [Vp + b(L + Lp)C/2]poiMi k{(L\ -f" CL2) Therefore, the dynamic-force-displacement relationship for a pressur¬ ized bearing is equivalent to a static spring stiffness with a lead or lag. When the bearing is deflected very slowly, equilibrium flows and pressure distributions are achieved in the clearance spaces, and the restoring force depends on the static stiffness kf. When the bearing is
Hydrostatic Bearings 199 deflected very rapidly, no appreciable flow can enter or leave the clearance space, and a restoring force occurs because of compression or expansion of the essentially trapped fluid film. The latter force depends on the compressibility of the film and is determined from the quantity k9(ri/r2). If a steady sinusoidal displacement of amplitude Y and frequency « is impressed on a bearing, then, in complex notation, the amplitude of the resulting force Fb is given by Fb = kj^1 + | Y 103T2 + 1 juT2 + If r2 > ri, as shown in Fig. 6-7, force FB lags displacement F. Then, since one component of the force is 180° out of phase with the velocity vector jo)Yf negative damping exists for any frequency w. Consequently, without external damping the system is unstable for r2 > ri. Conversely, if n > r2, the bearing is positively damped and stable. Thus the dy¬ namic spring constant in a stable bear¬ ing must be larger at high frequencies than at low frequencies. The magni¬ tude of n/r2 gives an indication of the amount of inherent positive damping in bearing. Effective negative damping Fig. 6-7. Vector diagram of force and displacement for compensated gas bearing. From Eq. (6-80a) and (6-806) the condition for stability is T1 r2 k(ai + a2) 1 2Vv + 1 > 1 (6-81) (6-82) 62 — 6i _6(L + L0)C Combining Eqs. (6-70) and (6-796) gives us KC = kib(L + Lp) /poA b2 - 6t p. 2 \p./ai-|-a2 The last two equations provide a basis for evaluating a bearing configura¬ tion required for stable operation. Since the value k{ai + a2)/(62 — 6i) is normally near unity, Eq. (6-81) indicates that the volume of any pocket bearing must not be of a higher order of magnitude than the clearance space volume and that the smaller the pool the more stable the bearing. If two bearings having an equal number of inlet regions, equal sizes, and equal values of inlet pressure ratio are considered, Eq. (6-82) gives (^•/pocfcet V I 1 _J_ ‘-‘P \ ^0 (k.) pocket 3 I i Xk.)i o. “2V 2/
200 Theory of Hydrodynamic Lubrication where IC stands for inherently compensated bearings. For practical bearing design Lp/L can vary from approximately one-half to zero. Thus the ratio above varies from 1.5 to approximately 2.25. From Eq. (6-81) (r 1 /T2) pocket (6-84) (n/rt)l0 S[2Vp/b(L + LP)C + 1] The pocket bearing is seen to have a higher spring constant, but in general it has lower stability properties. STEP THRUST BEARING Isothermal Operation 6-3. Compressible Lubrication. In a simple hydrostatic bearing, as shown in Fig. 6-8 the gas is introduced at the center of the pad and is distributed out at a radius Ro by means of grooves or reliefs. By Eq. (2-4) we have n _ h32irr dp {J ~ ~ From continuity under isothermal conditions we get pQ = p0Qo = const, which reduces the above equation to n Po _ 2trh3 dp Qo p ~ • 12m di By separating the variables and integrating, we get
Hydrostatic Bearings 201 Now when r = R, p = pa and therefore po \ TArpo R 0/ By rewriting and substituting Q0 from above, we get r - •" [1" tww' (|n r <°-85) or — T)2 — t71n2 — 71 2>i (r/flo) 0F PO P - (PO Pa ) ln (jR/flo) If we assume a constant pressure over the recess, we can then calculate the load W = TrpoRo2 — irpaR2 + J 2*r dr p 2t j pr dr = 2tt J p0 ^1 — Ciln^-^ r dr By using the probability integral Je~zt dx and by making the substitutions 1 - C, ln (r/fi.) = ^ = gl/C,6—x*/2 R0 d (jr^j = -e~lic’xe~x'n- dx it is possible to integrate the above expressions. The limits of integration are when -X- = 1 K o x “ Vcj_ “ when i = fo x = VA_21ni = ^ Therefore 2t I* pr dr = — 7rp0/2o2e2/Cl V^Ci J & x2e~x* d£ and, by integration by parts, IV - (1 - 6 jg + [*«-' - / «-• * The load is thus expressed in terms of the error function. (6-86)
202 Theory of Hydrodynamic Lvbrication Evaluation of Eq. (6-85) from R0 to R indicates that the pressure is nearly linear. Thus it is possible to simplify Eq. (6-86) by assuming linearity and integrating the pressure over the area over which it acts. We then have This equation shows that, when all other parameters are fixed, the hydro¬ static force is a function only of the entrance gauge pressure po — pa. Thus a change in entrance pressure is required for every change in applied load. To be practical, the pad must be self-regulating; i.e., the hydro¬ static force must increase with load. To satisfy this requirement, it is necessary to introduce a restrictor into the gas supply. Such a restrictor could be an orifice or a capillary. When an orifice is used to restrict the flow, the velocity of the gas upstream is negligible as compared to that at the throat of the orifice. The following relation is true: where po/p. is greater than critical, and where po/p. is less than critical; subscript c refers to critical pressure ratio. The use of two separate equations may be avoided if it is remem¬ bered that the expression (po/p«)1/n[l — (po/p«)(n_1)/n]H remains con¬ stant for all ratios of po/p* less than critical. Since the flow through the orifice must be equal to that through the pad, we get from continuity considerations: (po2 - pa*)*h* /2g fp»Yn Ti - /M(n-1)/TT 12ln (R/Ro) y/T~M 71-1) W L W J P = PO - Pa 0 < r < #o Ro < r < R and or (6-87) or h* =
Hydrostatic Bearings 203 6-4. Incompressible Lubrication. For viscous fluid flow in the flat- step bearing it is easily shown that the pressure distribution may be expressed by V - p. = (p. - p.) (6-89) The load-carrying capacity of the bearing then is given by TJ7 _ ^(PO “ Pa)(R2 — Ro2) 2 In (R/Ro) and the flow may be expressed by (6-90) o = (po - P°)*hz (6.on W 6m ln (fi/fto) ( ' Under the influence of centrifugal force and squeeze film, Eqs. (6-89), (6-90), and (6-91) become rMW + c. [„•-«.) + <*■ - «.■) **£] (6-89a) w - l<*’ - "•’> -«.[<*•+ *•■> - t^']) (6-90a) q ir^3 ["/_ \ i ft /pz p 2\ 6pr^h In (R/Rq)~\ Q - 6^mr7Ro) [(p0 ” Va) + C'(R ~Ro) v J (6-9 la) where C> =|(t + ¥) The total energy loss of the step bearing is made up of two parts: the pumping loss and the viscous friction. The pumping loss may be obtained by the use of Eqs. (6-89) and (6-91). Assuming a linear veloc¬ ity gradient, the viscous torque is dv dM* = 2irpcor3 ^ ri M* = Ifil (RA “ R°A) (6"92) There exists an optimum film thickness that will produce a minimum power loss, for both the friction and pumping losses are functions of film thickness. * Refers to friction torque.
204 Theory of Hydrodynamic Lubrication Adiabatic Operation If we consider adiabatic flow, the energy equation in addition to the momentum and continuity equations must be satisfied. A proper vis- cosity-temperature relationship must then be used, and for compressible fluids also an equation of state. By neglecting compressibility and inertia effects, making use of radial symmetry, and also neglecting velocities and velocity gradients in the radial and tangential direction in comparison to the changes of velocity in the y direction, we obtain for both compressible and incompressible flow the relations The energy equation, neglecting work of expansion and assuming constant temperature across the film, becomes dp _ yd2w Tr~ (0-93) The mass flow from continuity is 2rrpw dy By using the boundary conditions y — 0 w — 0 u — 0 y = h w = 0 u = ro) in the first two relations, we obtain for the velocity (6-94) (6-95) (6-96) By combining Eqs. (6-96) and (6-94), we get Trprh* dp ~ ~fyT dr (6-97) By combining Eqs. (6-93), (6-95), and (6-96) and integrating over an annular cross section, we get
Hydrostatic Bearings 205 Equations (6-97) and (6-98), together with a temperature-viscosity rela¬ tionship, must be solved for pressure and temperature. In addition, for compressible fluids the equation of state must be used: V = 9P&T 6-5. Incompressible Lubrication. From Eq. (6-98), it is evident that heat generated by tangential motion is much greater than that generated by radial motion. Therefore the energy equation reduces to ,, dT 2ro>Vr3 ,a qq\ gCv ~dr = —h— ^ ^ Assuming the viscosity-temperature relationship M = pve-w-To) (6-100) the temperature distribution becomes By substituting this expression into Eq. (6-100) and then into Eq. (6-97), we obtain the pressure as a function of radius: 6mo M . -4 p — Po = t r- In “•'('-InSM *•* [smw. <r‘" R,,) + *] (6-102) Next we want to evaluate M by using the boundary conditions p = pa, r = R: 6mo M . R4 Pa ~ Po = t—-—— v- In “'*(?nFS7 -■) + (6-103) M can be calculated from Eq. (6-103) and substituted into Eqs. (6-101) and (6-102) to give the temperature and pressure distribution as a func¬ tion of radius. The total load capacity is obtained by integrating the pressure over the disk: W = 27r [* rp dr + Tr(Ro2pQ — R2pa) J no
206 Theory of IIydrodynamie Lubrication This can be expressed as W = tR2(po — pa) + In ft — 2ft02 ln ft0 - (ft2 - fto2)(l + 2 ln ft,)] - - 'B [(ft2 •%/Ai + Ci) 4Ci2 VAi ln (ft2 y/Ai + Ci) - Jfto2 VAi + Ci) In (ft02 V^i + C.) + (ft2 V^4i - Ci) ln (ft2 - Ci) - (ft„2 VTi - Ci) ln (ft»2 - C,) - 2 -v/Al (ft2 - ft,2)] (6-104) The frictional torque may be found by integrating the shear stress over the surface of the rotating disk and is given by 6-6. Compressible Lubrication.. In the case of compressible lubrica¬ tion, by comparing the two right-hand terms of Eq. (6-98) for low and moderate velocities, it is seen that the second term is the larger. Conse¬ quently, the energy equation may be written as The variation of viscosity with temperature may be represented by By assuming that the pressure distribution under adiabatic conditions is similar to the isothermal case and varies nearly linearly with radius, we obtain dp = po - p. = K dr ft - ft0 By combining Eq. (6-106) with the above equation and integrating, we obtain By rewriting Eq. (6-101) and substituting the new equation for p and integrating, we obtain a more accurate expression for pressure distribu¬ tion than by assuming that it is isothermal. * Refers to frictional torque and not mass flow. where (6-105) (6-106) m = mo[1 + y(T — To)] (6-107) T - To y 1
Hydrostatic Bearings 207 If yDiRo2 < 1 the pressure distribution is given by P°2^ P* = (1 - yDiRo2) ln-^ + ^ (r2 - ft.2) - (yT„ — 1) |l — VI + yDt(r* - ft.2) + Vl - VW In ~ ?*>»«» + Vl + yDi(r + ft.)] r(Vl - y£>iR»2 + 1) I (6-109o) If yDiRo2 > 1 P°\~ pi = (1 - yDiRo2) In-g; + ^ + ««2) - (7T. - 1) [l - Vl + yDi(r2 - ft.2) + VyDiRo2 - 1 sec_I r yjyDiRo2'- 1 " V7£>.fto2 - 1 sec-1 ft. yjyDiyRito_ j ] where E = (6-1096) fiGiMgpo irh*y n ir/i’tf2 i-'i = Qp0Mgcv The frictional torque given by Jl/* = f*r3 Vl + Diyir2 - ft.2) dr n Jro integrates to jl/* = 15-^-g;j. ([■V»i(3ft2 + 2fto2) - 2][1 + 7/>i(ft2 - fto2)l« - (5/)i7fto2 - 2)} (6-110) SELF-EXCITED VIBRATIONS IN GAS-LUBRICATED STEP THRUST BEARING This section deals with the problem of self-excited vibration in step thrust bearings using compressible fluids. The stability analysis is based on a number of simplifying assumptions: 1. At equilibrium the recess area is subjected to a uniform pressure p0 and the pressure drop from the edge of the recess to the bearing periphery is linear. 2. For small deviations from the equilibrium point, this type of pres¬ sure distribution is preserved.
208 Theory of Hydrodynamic Lubrication 3. Changes in gas density are due primarily to pressure variations; therefore, the relationship p/p = g(RTo is considered to hold. 4. External damping may be neglected. 5. The motion is purely vertical. Based on the assumptions made and referring to Fig. 6-8, the pressure in the annulus can be written as P.. = Pr - (Pr - Pa) ^ T=T (6-111) it — /to The equation of motion using linear pressure gradients is m,A = 2r (J/ pr dr - J* p r dr) (6-112) r 2R3 - Ro3 - 3ft2/Sol r,, , = 3(R~Ro) J = ^ where /S.2 = /S2 3(R _ Rt) A€ = icRS mi = mass of the upper plate Referring to the Fig. 6-9, it is noted that the mass flow into the bearing depends on the recess pressure only, whereas the outflow is a function of the recess pressure, as well as the annular height. Fig. 6-9. Rates of change of mass flow about equilibrium point.
Hydrostatic Bearings 209 For small deviations from the equilibrium point (p and H) there are corresponding variations in inflow and outflow which, to a first degree of approximation, can be written respectively as (dM'\ Mn = Vd£) / = -“p ,, (dMA >(dMz\ (6'1U) Mn = \~w)ap + {~w)«H = 0P +eH where M is rate of mass flow and H = h — ho variation in film thickness. The time rate of change of the bearing gas mass content then becomes M = Mn - M22 = -(a + fi)p - BH (6-114) where a, 0, and 0 are all positive. The gas mass m2 contained between the bearing surfaces is m2 = 2x {J0*° (A + h)prr dr + f* hpar dr j (6-115) where A is the recess depth. By assuming pr = pr/g6iTo and making use of Eq. (6-111), we obtain pa If , x r - Ro ] p“ ff(R2’o g&To [pr (Pr ^ R - «oJ By substituting this into Eq. (6-115), integrating between the limits, and simplifying, we obtain m2 = -~~jr [hprAe + AprvRo2 + hpa{vR2 - Ae)] (6-116) The time rate of change of the bearing gas content M2 is evidently equal to the difference between inflow and outflow M and corresponds to the time rates of small deviations from the equilibrium point (p and H). ',+<w,7) where, by differentiation of Eq. (6-116)
210 Theory of Hydrodynamic Lubrication From Eqs. (6-114) and (6-117) we have qp + sH + (a + 0)p + 6H = 0 (6-119) and from Eq. (6-112) ^ - mi p~t.h p = (6-120) By elimination of p and p between Eqs. (6-119) and (6-120), the following differential equation is obtained: H + H + — H + —H = 0 (6-121) q m\q m\ q Equation (6-121) is of the form H + C2H + CXH + CoH = 0 (6-122) where all coefficients C are positive. Applying Routh’s stability criteria to Eq. (6-122), the following inequality must be satisfied in order to achieve stability CiCt > Co (6-123) Therefore a stability criterion is given by the inequality > - (6-124) $ s A study of Fig. 6-9 indicates that a high ratio of a + 0/0 corresponds to large values of the recess pressure po and small values of the annulus height ho (where the subscript 0 pertains to equilibrium conditions). For a given supply pressure p,f a favorable condition results if the maxi¬ mum possible load is being supported within the safety limits of a mini¬ mum annular height h0. Under those conditions the ratio of a/6 has a large value, though, unavoidably, 0 is small. Equation (6-118) shows that the value of q/s is proportional to the recess depth A and the annulus height ho and inversely proportional to the recess pressure p0. It can thus be noted that the values of po and h0 have an opposite effect on the magnitude of the ratios forming the two sides of the inequality (6-124). It is also apparent from Eq. (6-118) that the recess, which represents the bulk of the gas storage capacity, should have a minimum depth A in order to achieve stability. Referring to Fig. 6-10, it is evident that the magnitude of a depends on the manner in which the gas is supplied to the bearing. The three values of a shown correspond to three different conditions when gas is fed through a small nozzle «i, a larger nozzle a2, or a capillary ae. In each case, the load is the same, since the recess pressure p0 (but not the supply pressure p,) and annular height hQ remain
Hydrostatic Bearings 211 Recess pressure P Lorge-diom nozzle (2) Smoll-diom nozzle(1) Capillary (C) EE E ct2> a, > ac Fig. 6-10. Comparison of magnitude of the coefficient a. unchanged. The analysis shows that the recess depth has to be smallest with the capillary-fed bearing in order to achieve stability. Deeper recesses are possible for bearings with large orifices. For evaluation purposes of the recess depth the following two equations may be written: \R») [ p„r,(l - r22) J vl LW Jl I>(1 - (k -I- 1 ^r,(*+i)/Jt _ 9r,2/ifc“| \ + - n)»~J - V > A for ri > 0528 (f,-125) for (n < 0.528) (0-126) where a„ = cross-sectional area of nozzle „ /2gk(RTY 12m In (R/Ro) K = \T=t) 5? Ki = 0.532<RT» 12m ln
212 Theory of Hydrodynamic Lubrication SOURCES 1. Heinrich, G.: Uber Stromungslager, Maschinenbau u. Warmewirtsch., Wien, jg. 4, Heft 11, s. 176-179, 1949. 2. Heinrich, G.: Das zylindrische Stromungslager, Maschinenbau u. Warme¬ wirtsch., Wien, jg. 5, Heft 8, s. 136-143, 1950. 3. Heinrich, G.: Das Stromungs-Spurlager, Maschinenbau u. Warmewirtsch., Wien, jg. 6, s. 57-60, 78-87, 1951. 4. Heinrich, G.: The Aerodynamic Bearing, Part I: Laminar Flow, Maschinenbau u. Warmewirtsch., Wien, jg. 7, Heft 7, 8, s. 117-120, 129-135, 1952. 5. Richardson, H. H.: Static and Dynamic Characteristics of Compensated Gas Bearings, ASME Paper 57-A-138. 6. Weber, R. R.: The Analysis and Design of Hydrodynamic Gas Bearings, North Am. Rept. AL 699, 1949. 7. Hughes, W. F., and J. F. Osterle: Temperature Effects in Hydrodynamic Thrust Bearing Lubrication, ASME Paper 55-LUB-ll. 8. Licht, L., D. D. Fuller, and B. Sternlicht: Self-excited Vibrations of an Air- Lubricated Thrust Bearing, Trans. ASME, vol. 80, 1958. 9. Fuller, D. D.: Hydrostatic Lubrication, Machine Design, June-September, 1947.
CHAPTER 7 SQUEEZE FILM AND DYNAMIC LOADING DYNAMICALLY LOADED BEARINGS The preceding chapters have dealt with bearings subjected to a con¬ stant, unidirectional load. The parameters involved were considered to be independent of time and fixed with respect to a point in the bearing. The result was that for each set of conditions there was a corresponding position of journal or thrust runner which remained unchanged as long as the original conditions were maintained. However, in actual practice, it is likely that some of the parameters will be time-dependent functions and, as a result, the locus of the journal center will be some path other than a point. This applies particularly to the load which frequently varies in magnitude, direction, or both. The pressures in the fluid film and the position of journal center will thus undergo fluctuations com¬ patible with the variations in the applied load. Dynamically loaded bearings may be divided into two classifications: that of squeeze films and that of dynamic films. The first category refers to cases in which the journal does not rotate about its center; rather, the journal center moves, under an imposed load, along some path. If this motion is not cyclic, i.e., if the shaft center moves indefinitely toward larger eccentricities, we deal with the subject of falling bodies, which is concerned with the time it takes two surfaces separated by a fluid film to approach each other over a certain distance. It will be seen later that it is not necessary to have journal rotation in order to produce hydrodynamic forces in the lubricant. The other group refers to cases in which there is journal rotation about the journal center. The hydrodynamic forces here will consist of the contribution made by the progression of shaft center and that due to the rotation of the journal about its own center. It should not, however, be construed that these two components are algebraically additive; the resultant forces and displacements are of a much more complicated nature. Since the shaft may or may not rotate, and since the load may vary in both magnitude and direction, there are eight different conditions of dynamic loading, and these are listed in Table 7-1. In this table, as 213
214 Theory of Hydrodynamic Lubrication Table 7-1. Possible Conditions op Dynamic Loading U) Identi¬ fication letter up OIL Description w = 0; A 0 0 Constant unidirectional load shaft non¬ B X 0 Variable unidirectional load rotating C 0 X Constant rotating load D X X Variable rotating load (a ^ 0; E 0 0 Constant unidirectional load suddenly applied shaft F X 0 Variable unidirectional load rotating G 0 X Constant rotating load H X X Variable rotating load throughout the chapter, coP denotes the frequency of oscillation of a unidirectional load, while col is the frequency of rotation of the dynamic load. a> without a subscript is the rotational frequency of the journal. THE REYNOLDS EQUATION FOR DYNAMIC LOADING From Chap. 1 we have for the full form of the Reynolds equation + + + av‘ The right-hand side of Eq. (7-1) contains three terms; each of them con¬ tributes to the hydrodynamic forces in the bearing. The first term, 6 U dh/dx, represents the action of the journal rotating with a velocity U over a wedge-shaped fluid film given by h(x). In order for this term to generate positive pressures, it must be negative, since a wedge-shaped film implies that dh/dx < 0. The second term, 6hdU/dx, implies a variation of tangential velocity along the bearing surface, and in order that this term contribute to the positive pressures, dU/dx must be negative, i.e., the velocity must decrease along the fluid film. The last term is the expression for the velocity of shaft center and is responsible for the squeeze film action. Since \\ = dh/dt, it can be seen that, Fig. 7-1. Nomenclature for dynami¬ cally loaded bearings.
Squeeze Film and Dynamic Loading 215 when Vo acts in the same direction as the applied load, the film will decrease (dh/dl < 0) and the velocity will contribute to the load capacity. If the bearing is fixed, the shaft center will have instantaneous radial and tangential velocities. Referring these velocities to the vertical line <P = 0 of Fig. 7-1, they will be C de/dt, and e d(<p + <f>)/dt. Any point M on the surface of the shaft at angular position 6 will have tangential and normal velocities relative to M' on the surface of the bearing. These veloc¬ ities are made up of the components of the velocity of shaft center relative to the bearing center plus the velocity of the surface of the shaft Rio about its own center. Thus U = Rio + C ~ sin 6 — Ce 77-^- cos 0 at at r„ = C^coe« + C«^+*) sin* at at The infinitely long bearing is from above described by dx \u dx) dx dx By using the expressions for U and F0 from above and considering that (C/R) <K 2 and (e/R) « 2 cos $ and that 2Ce d{v + sin 6 = -2 ^±-*) ^ at at dd we obtain T,(‘‘IS) - «•*’ [(" - ,2~ - 2$) S + 20 J, “ *] By integrating between 0 and 2tt with the conditions that p(0) = p(2t) = p'a we obtain _ a (R\2 | 2 + e cos Q ( 0 0 d<f>\ e . Q p p0-6/z^cj j(l € C()S 2^ - dt) 2 + €2 .fir —1 1— 1—) (7-2) ^ + € COS e)2 (l+€)2Jd*j K ' where iol = d<p/dt is the frequency of rotation of the applied load. In the expression above, the term (ioL + d<f>/dt)e sin 6/(2 + e2) represents the contribution of the tangential motion of shaft center, the terms in de/dt its radial motion. If ioL = d<t>/dt = de/dt = 0, Eq. (7-2) reduces to the Sommerfeld solution.
216 Theory of Hydrodynamic Lubrication Equation (7-2) integrated along and at right angles to the line of centers yields Sin* - * X( 2«L-2d-£) (7-3«) 12irsS (2 + f2)(l - a \ L dt COS 0 ( 1 de (7-3 b) 12t2S (1 - €2)^ 0) dt where Eq. (7-3a) is the load capacity due to the wedge action and Eq. (7-36) is the load capacity due to the squeeze film. This set of equations provides a relation between e and <t>, the phase angle between load and line of centers. In these equations, £, a>, and can be functions of time. This may make the integration of Eqs. (7-3) difficult but when accom¬ plished these equations would, through the relation between e and <f>, yield the cyclic locus of shaft center as well as the instantaneous resultant hydrodynamic force. Through a relation between the applied load and the minimum film thickness of the shafts orbit, a comparison with the performance of steady-state bearings is possible. By using the previous expressions for the radial and tangential velocity in the equation for the infinitely short bearing which is given by Zn/TTdh .udU OT/\/2 L2\ p~p‘ = ¥\uTx + hfc+2Vj\z ~T) we have, by again neglecting terms of order higher than C/R [i (“ -**L-2 sin 6 - § cos *] (7-4) where pa is the ambient pressure. By integrating over 2t with p( 0) = p(2ir) = v(±^J = p* we have ~ (d) 2(1 -«*)*«(“ 2wt 2 dt) (7'5o) cos 0 /L\2 1 -f 2e2 1 de (n cn 4^S -(d) (1 (7-56) From Eq. (3-22) the instantaneous friction coefficients are R f , € . 2ir2S Cf= ±2 S1P»+(1 _.»)» (7-8) and the average power consumption over a period of h — t\ is 2 R2L f1'
Squeeze Film and Dynamic Loading 217 Equations (7-3) for the infinitely long and Eqs. (7-5) for the infinitely short bearing give a relation between the applied load and motion of shaft center. The solution of these equations presents no problem when the motion of the shaft center de/dt and d<j>/dt is known and the quantity to be calculated is the resultant load. However, in practice, the quantity that is usually known is the load, and the motion of the shaft center and the resulting eccentricities have to be calculated by solving the set of differential equations in </> and e. CYCLIC SQUEEZE FILMS IN JOURNAL BEARINGS For the type of motion labeled A and B in Table 7-1 the shaft does not rotate, the load vector is fixed in space, and the shaft center follows a path determined by the nature of the imposed load. Since here O) = COL = 0 we have from Eqs. (7-3) 1 P sin <f> 2e d<f> 6t(jB/0*m (2 + e2)(l - e2)* dt 1 P cos <f> 1 de for {R/CYii ~ (1 - c2)** dt By dividing the two expressions by each other, (2 + e2) de (7-8a) (7-8 b) [ d<f> = fi J tan <*> J 2«(1 + €2; and by integrating, we have sin <t> = k'(1 ~ <2)* (7-9) This equation describes the locus of the shaft center under a unidirec¬ tional load. It will be seen that the locus depends on the constant of integration fc', which is determined by the initial position of the journal. The various loci as given by Eq. (7-9) are shown in Fig. 7-2a. The particular path followed by the journal depends on its position at the instant the load was applied. The journal center once located on a given path will always remain there regardless of the magnitude and nature of the load. This can be seen from Eq. (7-9), which determines the locus without containing the expression for P. The type of load will only determine how far and how fast the journal center will move along this fixed path. To determine the relation between load and eccentricity, either Eq. (7-8a) or (7-8b) can be used by eliminating <f> between the two equations.
218 Theory of Hydrodynamic Lubrication io) Fig. 7-2. Locus of a non rotating journal under unidirectional load: (a) infinitely long bearing; (6) infinitely short bearing. For an infinitely short bearing the two equations equivalent to Eqs. (7-8a) and (7-86) are P sin </> 7re d<f> (L/C)V 2 (1 - e2)* dt P cos </> r 1 + 2e2 de (L/C)v “ 2 (1 - e2)» It and the locus of shaft center is given by sin $ = k — (7-10a) (7-106) (7-11) The characteristics of this locus are similar to those of Eq. (7-9), and these are given in Fig. 7-26. The motion of shaft center is, by use of Eqs. (7-10), given by + (7‘l2) k cot 0 = 7- (^j J l>dt + K' where k and Kf are constants. Returning to the infinitely long solutions, we have for the particular case of a shaft traveling along the vertical center line (<f> = 0) from Eq. (7-86) d - ‘ ..n- - 7"j (‘ ..,n. = / 777 !.,,-, [ Pdt (7-13) - e~) (1 - erp- for (R, ( )’n J
Squeeze Film and Dynamic Loading 219 This expression provides a relation between load and eccentricity, and its exact form depends on P. 7-1. Constant Loads (Type A of Table 7-1). If P = const, Eq. (7-13) yields (i - «*)» “ (i - «i2)» = &T(ft/C)V(< “ ll) (7'14) Equation (7-14) determines the time it takes a journal to drop a distance e — €i when subjected to a constant load P. This subject is discussed in more detail in the section on falling bodies. 7-2. Alternating Loads (Type B of Table 7-1). If the load is given by P = Po sin copt where Po is the maximum amplitude and wj> the frequency of oscillation, and using c = €i at (apt = r/2, we have from Eq. (7-13) € <i cos (apt (n . - (7-15) (1 - e2)* (1 - ei)* \2ttW where *s' = (?)2£f„ Equation (7-15) describes a periodic motion which is not neces¬ sarily symmetrical about ei. If «i = 0, the oscillation is symmetrical about the bearing center and the maximum eccentricity is given by setting cos caPt = 1 or -™x = - — (7-16) (1 - \2t2S' 0) The displacement as seen from Eq. (7-15) is always lagging the load vector by x/2. For square-wave loading P = +Po for the first half cycle w/(aP P = —Po for the second half cycle Since the displacement lags the load vector by tt/2, the maximum eccen¬ tricity will occur one-fourth cycle later, i.e., at tt/2(aP. By use of «i = 0 at t = 0, Eq. (7-15) yields at t = t/2cop ^max 1 (1 - *L.)M 24ir<S'' (7‘17) 7-3. Rotating Loads (Type C of Table 7-1). In a symmetrical bearing with the load constant and rotating at a uniform frequency, it is reason¬ able to expect a fixed relation between load vector and displacement as
220 Theory of Hydrodynamic Lubrication well as a constant eccentricity. Although this cannot be proved a priori, it is assumed here that for this type of motion _ d# _ de _ n dt dt and thus by Eq. (7-3) P sin <t> € , 6r(ff/C)V (2 + 0(1 - €2)* ^ L) P COS <f> \2ir\R/C¥n These two equations yield = 0 P.-12,, (*)■ „ (2 + ,,;(l . ,,)W (MS) Equation (7-18) gives the load carried by a nonrotating journal subject to a constant load rotating at an angular velocity wz,. This expression is similar to the expression for a steady load carried by a journal rotating at w = except for the sign, which indicates that the two resultants are always 180° out of phase. When the equations for the infinitely short bearing are used, the load capacity is (*)’©' <r^r» <™» and <f> = t/2. The comments made about the infinitely long solution apply also to this last expression. NONCYCLIC SQUEEZE FILMS This section is concerned with the time it takes two bodies separated by a thin fluid film to approach each other over a certain distance when one of them is subjected to a velocity V. The velocity V is usually imparted by a load W which will be considered throughout this treat¬ ment as constant in magnitude and direction and symmetrically placed with respect to the boundaries of the system. One such expression is given by Eq. (7-14), which determines the time it takes a journal in a full sleeve bearing to travel the distance e — «i. This expression was developed from the basic equations, (7-8). The various geometries con¬ sidered here can be analyzed much more simply by equating the flow as predicted by Eqs. (1-13) to that which results from the displacement of the configurations involved.
Squeeze Film and Dynamic Loading 221 7-4. Journal Bearings. From Chap. 1 for JJ = 0 we have D _ Lhz dp Q ~ 12nR dd From Fig. 7-3 the flow at any 6 due to the velocity V is Q = VLR sin 6 The two equations yield p = (c/R)'R [>(1 -«cos ey + Cl] (7'20) When this p is integrated for the case of a full bearing and V is replaced by the expression V = C de/dt, we obtain W — 12t/xL „d* (7-21) W ~ (C/R)\ 1 - e2)^ ° dt U Zl) which when integrated again yields _ _12irpLR T €2 _ €l 1 (7-22) ” (C/R)*W [(1 - <22)* (1 - U ' the same expression as Eq. (7-14). If Eq. (7-20) is applied to a 180° partial bearing, then the boundary conditions are p(±x/2) = 0, which yield H (C/R)*Re This expression integrated between 0 and ir yields W = 2LR (' p cos 6dd (7-2S> / v 1
222 Theory of Hydrodynamic Lubrication By replacing V by C de/dt and integrating, we obtain * j 24uLR T €2 / 1 + e*Y4 1 A + €i\^l l(i-'«,«)>*tan v-«iV (i-*i2)“ v-«i/J (7-25) 7-5. Spherical Bearings. The configuration considered here is that of a hemispherical seat which, owing to symmetry, has a cross section identical to that of Fig. 7-3. For a sphere the amount of fluid passing a conical element of surface is according to Eqs. (1-13) Q _ (2tcR sin 0)C3(1 — e cos 0)3 dp Q 12pR dd The displacement of the fluid at any point in the seat is Q = tVR2 sin2 0 The two equations yield a relationship , _ 6nV sin 0 dd V ~ (C/R)*R( 1 - € cos 0)3 which when integrated with p(±ir/2) = 0 gives.us v = {c/R)*Rt [(i - ecos ey ~ *] (7_26) The equation for load capacity of a hemisphere is W = 2tR2 f;/2 p sin 0 cos 0 dd W = WfRY [i3 ln (1 _ €) + ^0^1) ” 2e] (7‘27) This equation integrated for e and t gives 3*nR2 r«2-€l i+tl2. . N 1 +«*21_ si (C/R,yw[ t,ei + «,2 ( l} (22 ( €2)J (7-28) By similar methods it can be shown that for a spherical seat of a span equal to 2a the time-eccentricity relationship is given by from which A t= tm puw f (l + -V)ln 1 f1 + —2Vn rr — (C/RyW \\ c22/ 1 — €2 cos a \ d2/ 1 + ei cos a sin2 a sin2 a (1 l\ 1
Squeeze Film and Dynamic Loading 223 7-6. Conical Seats. By methods similar to those used above, the full cone of Fig. 7-4 (a) has a load capacity of W = (7-30) 2hz sin4 a and the time-eccentricity relation is At = (A - A) (7-31) 4 IF sin4 a \h22 h\l) For the truncated cone whose smaller radius is R\ the two equations are W = SttvlV At = 2h3 sin 3 Try. *a[Ri Rl ln («,/«,) J 4IP sin4 a R2 R i4 In (R2/Ri) (R22 - ln (R2/R 0 J W hx*) (7-32) (7-33) (4) Fig. 7-4. Full and truncated cones: (a) full cone; (6) truncated cone. 7-7. Sliders and Rectangular Plates. The rectangular plate poses a two-dimensional problem, and we must return to the basic differential equation of viscous flow for a solution. We can think of falling plates in terms of a slider with no tangential velocity and a normal velocity equal to V. If we apply the Reynolds equation to our case, we have for h = const d2p d2p 12m v dx2^ dz2 hz (7-34) which is recognized as the Poisson equation. Let us assume the solution of Eq. (7-34) to be given by oo p(-v) = ( ^ a. cos ^)z(z) n = 1,3,5 where Z(z) is an unknown function of z alone and An is a constant to be
224 Theory of Hydrodynamic Lubrication determined. Then after substituting this solution into Eq. (7-34), we obtain oo oo 21 nxY a nrx ?{ \ i X' a nTX d2Z(z) 12/u T/ (b) A» cos T Z(-z) + L, A" cos ~B~ W = T>- F n-1,3,5 1,3,5 Since i- j mrx - cos n-773.5 we can write oo oo (mr\ . rvKX rjf , . . mcx d2Z{z) - b vb) a-™~bz^ + 2/ A*cos-B^y IOC IOC 00 4^2 -J- n-1,3,5 n-1,3,5 4 12m F 1,3,5 which is satisfied if . <PZb) /t»Y . _48„V(-1)<-W* “5?“ _ U) a"z(z) ~ TV —n— (7'35) Expression (7-35) is an ordinary linear differential equation with a homo¬ geneous and particular solution given by rjf \ D • U nirZ . n 48mB2F (—l)(n-1)/2 « ocv Z(z) — Bn sinh --g |- Cn cosh ^ 7r3/i3 nzA (7-36) Thus the general solution to Eq. (7-34) is p(x,z) = ^ An cos ^ sinh ™ + Cn cosh n-1,3,5 AQ.. D2T7 ( 1\r«-l)/2l (7-37) 48mB2F (-1)(»-d/*J irzhz nzA One of the terms above can be eliminated by the requirements of sym¬ metry in the z direction, or Bn = 0. From the boundary conditions p = 0 at z = ± ^ we must have for all x n , mcL 48MR2F(-l)(n-1)/2 n Cn cosh 2B ^hl n,A^ 0r C = — (— l)(n-l)/2 \ ° " irzhznzAn K } cosh (mrL/2B)
Squeeze Film and Dynamic Loading 225 which when substituted into Eq. (7-37) and simplified yields for the pres¬ sure distribution v(x£, _48mB»P V (-!)<-)/»[ cosh(n«/g) P(’ ’ T*h> Li n3 [cosh (nwL/2B) J B n-1,3,5 (7-38) The total load by integrating Eq. (7-38) yields w=2ii p{x’z) dx dz=2 tanh § - £) n-1,3,5 (7-39) By writing F = d/i/cft and integrating Eq. (7-39) for a relation between 2 and h, we have 2 (£-£-*«)] <M»> n-1,3,5 7-8. Elliptical and Circular Plates. The differential equation (7-34) used above applies also to any other arbitrary configuration. For an elliptical plate the boundary conditions are x2 z2 p = 0 on - + - = 1 and we thus have for an immediate solution p(x,z) = Cl(g + g- i) (7-41) where a, b are the major and minor semiaxes of the ellipse. Equation (7-41) satisfies the boundary conditions, and all that remains is to satisfy the nonhomogeneous terms of Eq. (7-34). By using Eq. (7-41) in the Poisson equation, we have or (l + 1\- 12"F 2ti(? + 5y ~~v~ from which r 6 uVa2b2 1 h3(a* + b*) Thus the pressure distribution is
226 Theory of Hydrodynamic Lubrication The load capacity is, by integrating Eq. (7-42) b r ay/\ — t'/b* (7-43) By writing V = -rr and integrating, we have ti - h = 2 (a2 + b2)W 3TMa363 (7-44) Circular plates are a special case of the above expressions and can be calculated by setting a = b = r. Thus, for the h-t relation we get 7-9. Miscellaneous Configurations. By methods similar to those described above or by transplanting analogous solutions from problems in torsion where the differential equation is also that of Poisson, results for other shapes of falling plates can be extracted. Some of these are given below. 1. Circular rings of inner radius ri and outer radius r2: By setting r\ = 0 in Eqs. (7-47) and (7-48), the solution of a full sector of radius r is readily obtained. h - 2. Circular sector of angle /S and bounded by radii r\ and r2: n-1,3,5 n0 [4 — (nx//3)2]2 r22nT//3 — ri2nr//5 2 x 1 (r2 2+nWfl — ri2+nWfl) ] (7-47) 4»2(2 - nvl&y r24 - r,4 (7-48)
Squeeze Film and Dynamic Loading 227 3. A circle of radius a, a portion of which has been cut out by another full circle of radius 6, as shown in Fig. 7-5: If b = 0, the above equations reduce to those given for a full circle. 7-10. Constant Unidirectional Loads. In steady-state hydrodynamics a constant nonrotating load yields journal attitudes which are fixed with respect to the bearing. However, if the more general approach of an arbitrary initial shaft position is taken, then the resultant solution is one of the shaft center moving in a cyclic orbit. The locus will depend on the applied load and initial shaft position and will have as its pole the steady-state eccentricity corresponding to the applied load. Physically, this shaft behavior corresponds to the free oscillation of an undamped system in which the shaft, under the impact of a suddenly applied load, first overshoots its proper attitude and then is forced back to its original position by the excessive hydrodynamic forces resulting from an eccentricity which is too high for the applied load. Thus the journal, although subject to a constant unidirectional load, will continue in an orbit around its theoretical eccentricity. Only if the initial shaft position corresponds to the magnitude of the applied load will the orbit be a point. The above phenomenon is a result of a hypothetical system devoid of all damping. In practice, damping will soon bring the shaft to a steady-state position in accordance with the solutions of Chap. 3. However, these undamped oscillations play a part in the various forms of hydrodynamic instability discussed in Chap. 8. Fig. 7-5. Section of a circular plate. DYNAMIC LOADING OF JOURNAL BEARINGS
228 Theory of Hydrodynamic Lubrication This case, therefore, corresponds to condition E in Table 7-1, and from Eqs. (7-3) by setting o>l = 0, we have (2 + 0(1 / 2 d<t>\ _ sin <f> - €2)*\ 0)dt) ~ 127r2£ 1 de/dt __ cos <t> (1 - €2)* 12*2S With d<f>/dt = de/dt = 0, these two equations yield the steady Sommer¬ feld solutions e° 1 J - T (7-51) (2 + «o2)(l - to2) 12t2S TC 2 where e0 is now the pole of the general orbit traveled by the shaft center. This orbit is obtained by eliminating dt between the two equations above J'* sconst S = 0.0214 it) Fig. 7-6. Journal locus under a constant load suddenly applied: (a) infinitely long hearing; (6) infinitely short bearing. and integrating the resulting expression in e and <t>. The result of the integration is sin <f> = 12* 2S 5«(1 - €2)” + K (1 - €2)** (7-52) which gives the cyclic locus of shaft center. This locus depends on the value of the constant K, which is determined by the initial position of shaft center. Equation (7-52) represents a family of curves with one extreme being a point locus as given by Eq. (7-51) and the other a circle of radius C corresponding to an initial shaft position of e = 1. A sample solution for the case of a pole of e0 = 0.7 is given in Fig. 7-6a.
Squeeze Film and Dynamic Loading 229 The periodicity of this motion is given from the starting equations by 1 — d<t>/dt/u. The ratio of vibration frequency to the rotational speed is d<f>/dt 1 T -i s*n </>/12t2£ ] m KO>k co ~ 2 L c/(2 + e2)(l - e2)^ J {7'b6) which is seen to depend on the instantaneous position of the shaft center as well as on the pole of the orbit. When this expression is examined for the particular case of e = 1, the average value of (d<f>/dt)/co is always one-half. For all orbits that have c < 1 but enclose the bearing center and for all very small orbits, this ratio is less than one-half. Thus, in general, the frequency of shaft oscillation is always slightly below one-half of shaft speed. For an infinitely short bearing, the same conditions yield from Eqs. (7-5) 2(1 - «*)» (* ~ a J 4r*(L/D)!S 1 + 2e2 de/dt cos (1 + €2)W ~u~ “ 4t2(L/D)2S By elimination of dt between the two equations and integration, (7-54) This yields a family of curves similar to that obtained from the infinitely long solution; an example is given in Fig. 7-6b. By using the expression of Eq. (7-54), we can obtain from the preceding equation the motion of shaft center in terms of e and K as a parameter. This yields d<f>/dt\ __ sin <f> 2’‘ (d)‘ «■ <riW. - [:1 + 4',K'S’ (s)T “ - (t - to) 2 + |^1 + 2ir*K2S2 (^ (7-55) The frequency of shaft motion is from the term sin At seen to equal or exactly one-half of journal rotation. The pole for each orbit is given by setting d<j>/dt = de/dt = 0 or S-2'*(b)’(T^7)» W 7-11. Variable Unidirectional Loads. For the case in which the fre¬ quency of both the applied load and journal rotation may vary let us define by
‘230 Theory of Hydrodynamic Lubrication a — angular displacement of journal \p — angular displacement of load P = PoP(wpt) = PaP(r), where P0 is the maximum load amplitude and o)p its frequency f = (! J = tan (sin_1 f) 0" (cy i\ From the above, w = da/dt and cop = d\p/dt. By assuming that is in most cases related to w/> by a multiple of its frequency, we have from Eqs. (7-3) f ~t~ t y ^p ^ r / \ \ oil P(r) sin <f> 2 + 3f2 f « rfr ^t) 2^t) ( ) 03p dt _ P(t) cos <£ o) dr 12tt2*S’o (7-57 b) A plot of [(1 + T2)/(2 + 3{*2)]$* vs. t will show that, beginning with t > 0.7 or € > 0.58, the expression [(1 + f2)/(2 + 3f2)]f can be approxi¬ mated by a straight line, the deviation there being no more than 1% per cent. Thus, except for lightly loaded bearings, [(1 + $*2)/(2 + St2)]t can be replaced by at 4- 6, where a and b are constants. Writing now .r = 12tt25o — (at + b) (j) we have from Eqs. (7-57) X (a - - 2*) = P(t) sin « - ~ = P(t) cos <t> ar a dr These last equations give x (that is, e) and <f> as functions of r = upt. For our case the load is unidirectional, thus d\p/dt = 0, and the equations become - *^) = P(T) sin * (7'58a) - ~ = p(r) cos 4> (7-586) a ar The solutions of these equations will, of course, depend on the form of loading P(r). In the equation above the value of which should be used as a basis of comparison of dynamic and steady-state loading is given by
Squeeze Film and Dynamic Loading 231 A similar approach to the infinitely short bearing yields equations corresponding to those of Eqs. (7-57) in the form {wp d . . . . . „ . _ P(t) sin <f> 2 u dr l“(T) “^(t) 01 4ir2(L/ £>)2So cop d£ _ P(t) COS co dr 47T2(L/Z))2£o (7-60o) (7-606) where £ = t/{\ — e2)*-. If we write the two equations become x' — (<* — 2\j/ — 2<f>) = P(r) sin <f> 2 = P(r) cos <f> ar ar and with d\p/dt = 0 •c'(J-2S) = PWsin^ (7-61“) dr' 2 —■ = p(r) cos 0 (7-616) Equations (7-61) have the advantage of not containing the approxi¬ mation of a linearized £ which restricts the infinitely long solution to moderate and heavy loads (c > 0.55). We shall now use the developed relations to examine some of the more common modes of variable uni¬ directional loading. Sinusoidal Loading P(r) = sin a)pt If the journal speed is constant, we have from Eqs. (7-58) da/dr = co/cop, and thus x — 2 = sin r sin <f> (7-62a) ~~r~ s*n r C()S ^ (7-626) a dr These equations cannot be easily solved; they contain two integration constants which, because of the cyclic nature of loading, must be so chosen that the path closes after a complete cycle in r. The solutions must be obtained by numerical methods involving a trial-and-error pro¬ cedure. The calculations are summarized in Eig. 7-7, where the ordinate gives the ratio of xm%x and thus cmmx of a sinusoidally loaded bearing to that of a steady-state bearing in terms of the frequency ratio cop/w. The
232 Theory of Hydrodynamic Lubrication 7 6 J5 n 4 SI3 z / z z 0 0.5 1.0 2.0 Vplu 3.0 4.0 Fig. 7-7. Load capacity for sinusoidal loads relative to constant loading. curve has not been calculated for values up/u < and its behavior there is uncertain. At w?/« = 0 the ratio is simply the steady-state solution. At up/w = the load capacity is zero. As the load frequency up rises, the load capacity also rises, and as up/u—» oo the ratio (up/u)/xm„ approaches the line 3up/u. The locus of shaft center changes its form depending upon the frequency [o) (6) (0 Fig. 7-8. Journal locus for sinusoidal loading: (a) infinitely long bearing; (b) infinitely short bearing; (c) general appearance of journal locus under sinusoidal loading.
Squeeze Film and Dynamic Loading 233 ratio cop/co. If this value is below one-half, the loci will assume the form of the first diagram of Fig. 7-8a with the orbits becoming flatter as cop/co —> 0. Above wp/w = the orbits will assume the shape of Fig. 7-86. Figure 7-8c shows the various possible loci on a single plot, and Fig. 7-9 shows two specific examples of the load capacity of journal bearings subject to a sinusoidal load. Square-wave Loading. In this type of loading Hr) = +1 P(r) = -1 for 0 < r < t for t < r < 2t The solutions here are the sum of two solutions for a fixed load acting in opposite directions over the two respective half circles. By denoting by «m the eccentricity corresponding to either </> = 0 or <f> = t, we have from Eq. (7-54) Fig. 7-9. Maximum eccentricity for sinusoidal loading. D (1 - O* = K (7-63) This K substituted in Eq. (7-55) gives the time at which the shaft center is at these points, namely, the two values of t resulting from t — to — - sin-1 - co -1 1 + 4t2S02 (-Y—* \DJ (1 - The difference between the two roots is the time required for the shaft center to complete one-half cycle which in turn must equal one-half the load cycle. This, therefore, gives a relation between em and cop: tan^ = 2^o(^),(r^ (7-64) Figure 7-10 shows sample solutions using the equations of both the infinitely long and infinitely short bearing. It is seen that, when co > J^cop, this em is also the maximum eccen¬ tricity; however, when co < the maximum eccentricity occurs at right angles to em. This can be found, then, by setting sin <f> = ± 1 in Eq. (7-54) 1 (1 - eJY (I - eLJ3 t*Sq(L/D)* (1 - €L,)* (7-65)
234 Theory of Hydrodynamic Lubrication (o) w/»/w = 1/4 u)p/u) = 3/8 U) Fig. 7-10. Journal locus for square-wave loads: (a) infinitely long bearing; (b) infinitely short bearing.
Squeeze Film and Dynamic Loading 235 Thus, by combining the relation between coP and €m of Eq. (7-64) with we obtain the relations between cm and 'LV 2v2 (1 - ( ,T ^max W ° (1 - i T “ pr* = 1 “ sec 3 — )'■ 4 up 7T O) = tan- — )92 4 w/> \ < i (7-66a) 4 a> Z i < — (7-666) Z a) By comparing with Eq. (7-56), it is seen that for a constant load the quantity up ^ 0 1 w~ 2**S0 (q) (1 jTjJ* (7-67) equals unity. This ratio thus represents the relative load capacity referred to the hmla of dynamic and steady loading. This ratio is plotted 8 7 6 ?5 ■*4 3 2 1 CL -L > Squ uie wuve iuuc p°\ i \ / 0 0.5 1.0 2.0 dip/d) io) 3.0 4.0 ^ 6 S 5 ~ E r4 —r3 . | ^ 2 ~— 1 £ °0 0.5 1.0 ^Squore-wove load t 1 1 > 2.0 uip ha (b) 3.0 4.0 Fig. 7-11. Load capacity for square-wave loading relative to constant loading: (a) infinitely long bearing; (6) infinitely short bearing. in Fig. 7-11. The plot is not valid below u/up < 0.2; at high values of wp/w, it approaches the value of (4/t)(o>p/w). A comparison of the relative load capacity of bearings subjected to various forms of an alternating load is given in Table 7-2. Table 7-2 <t)p/u) Form of load 0 0.5 1 2 3 F =P0 1 0 1 3 5 P = Po sin (apt 1 0 1.74 4.72 7.90 II H- 1 0 1/25 3.27 5.24 7-12. Constant Rotating Loads. For a constant load rotating with a frequency wl and by assuming the phase and amplitude of the orbit of
236 Theory of Hydrodynamic Lubrication shaft center to be constant, we have from Eqs. (7-3) (2 + (2)(1 - e2)* (* ~ 2 = ± ('7'68) <t> = ±x/2 Equation (7-68) can also be obtained by adding algebraically the load capacity for a nonrotating journal with a load rotating at a frequency ujl as given in Eq. (7-18) to the load capacity of a journal rotating at w with a nonrotating constant load as given by Eq. (7-51). The two load capacities are 180° out of phase, and an algebraic addition yields the resultant load capacity. The resultant of Eq. (7-68) is seen to depend on the value of 1 — 2col/w, being 1 at o>z,/a> = 0 (steady-state solution), becoming zero at wl = and then rising continuously with a further rise in oil. The special case of a)L = when the shaft center will progress in an orbit with no resultant hydrodynamic force is the theoretical back¬ ground for the experimental fact that a very lightly loaded journal is subject to severe vibration whose frequency is about one-half that of the journal speed. The progression of the shaft center is in the same direc¬ tion as the rotation of the journal. The phase angle between the load and line of center is theoretically tt/2. To determine the nature of the orbit of vibration, we can for a moment discard the assumption of de/dt = d<f>/dt = 0. Then from Eqs. (7-3) we have sin <f> (2 + *2)(1 - t2)* V w " dt) 1 1 de cos <f> (1 - €2)* a) dt ~ 12ir2£ and, by eliminating dt between the two equations and integrating, we obtain . , 12x2S(l - 2«l/«) , ^ (1 - €2)*4 Sln * = 5t(l — t2)** + K — (7'69) This last equation is similar in form to Eq. (7-52) except for the term 1 — 2a)l/o). The orbits thus will be identical with those of Fig. 7-6 except that they will be related to the line of centers, which is itself in motion. The orbits relative to the bearing will be quite complicated and will in general not be a closed path. For the infinitely short bearing a similar treatment yields »in . , _ K, » - <■)“ Both the infinitely short and infinitely long solutions show that the line of centers is always at right angles to the load and that the eccentricity
Squeeze Film and Dynamic Loading 237 Fig. 7-12. Dynamic loads in a connecting-rod bearing; numbers indicate crank angle in degrees /map = 109.5 psi, Wm»x = 9,360 lb, N = 2,000 rpm, two-stroke cycle. (After Burwell.1) of the shaft is the same as if the load were fixed provided its magnitude is modified by the factor (1 — 2u3L/o)). 7-13. Variable Rotating Loads. It can be anticipated from the diffi¬ culties with the simpler modes treated above that the case of a load varying in both magnitude and direction presents a formidable problem. The general expression for this type of motion is given by Eqs. (7-57), where P(r) and ^(r) are input variables. These cannot in most instances
238 Theory of Hydrodynamic Lubrication be expressed mathematically. Even if that were possible, the resulting differential equations would be extremely challenging. The difficulty of the problem can be seen from Fig. 7-12, which shows the polar load diagram for the bearing of a connecting rod in a two-cycle Diesel engine, with its corresponding locus of shaft center as obtained by numerical calculations. Even if the load diagram could be mathematically stated, it is doubtful it would be possible to represent analytically a motion as complex as that obtained here. Practical problems of this nature call for an individual treatment of each case. DYNAMIC LOADING OF JOURNAL BEARINGS WITH NO NEGATIVE PRESSURES Equation (7-2) was integrated over the entire region 2?r and, as in the case of the full Sommerfeld solution, it includes negative pressures of a Fig. 7-13. Nomenclature for analysis excluding negative pressures. magnitude that cannot possibly occur. We shall now integrate the equation for dynamic loading only over the positive pressure regions by setting the pressure in the remaining portions equal to zero. This involves finding the zero points, i.e., the angles at which the pressure wave begins and ends. This analysis is taken from Ref. 3. We may start with Eq. (1-11) and writing Vo = dh/dt with the nomenclature as given in Fig. 7-13. By writing for the film thickness h = C[ 1 — c cos (\p — 7)]
^ • • Squeeze Film and Dynamic Loading 239 we have (C[R]_ jjj _ t cos ^ fL? _|_ 3[J _ t cos _ 7)]2e s;n _ y) 6/uo> ( d\J/2 d\f/ - e sin - 7) - ^ [- cos - 7)] ^ - € sin - 7) ^ =0 By use of _ p(C/R)2 F=*dl r = 2rfy /XO) J 03 dt 03 dt the equation above can be rewritten as — 7) - g(l - 0) [1 — e cos (^ — 1 = 0 (7-72) d2n 3c sin — 7) dn „ E cos — 7) — c(l — G) sin (\J/ — 7) d\p2 1 — c cos (\J/ — 7) d\p [1 — e cos (\p — 7)]3 By using the substitutions \j/ — 7 = 0 n = and noting that (1 — c cos 0)2 an _ an a2n _ a2n e+ “ de du2 “ de2 we have in terms of the new variables u and 0 d2U du (1 — € cos 0) -rTz — e sin 0 — — 2c(cos 6)u = 6[c(l — G) sin 0 — E cos 0J dd dd (7-73) Assume a solution of the form u(d) = Ci sin 0 + C2 sin 20 + C3 cos 0 -f- C4 cos 20 By replacing w, w', and w" in Eq. (7-73) and equating the coefficients of like terms on the right- and left-hand sides of Eq. (7-73), we evaluate the constants to give 12«(1 - G) . a . 3c2(1 - G) sin 20 . VIE UW = 2+«» Sm 9 + 2~+«» + C0S 9 ME oa cos 20 2 -f- € or, by replacing u(d) by its original expression, n = p(c/R)* = 1 [-12«(1 - G) sin e no) (1 — c cos 0)2(2 + c2) + 3c2(l - G) sin 20 + 12# cos 0 - 3eE cos 20] (7-74) The constant of integration in obtaining Eq. (7-74) was set equal to zero on the requirement that p(0) = p(7r) =0 for the steady-state case.
240 Theory of Hydrodynamic Lubrication As was stated in the beginning, the pressures are to be integrated only over the positive region. This requires the knowledge of the zero points of Eq. (7-74), i.e., the angles 0i and 02 at which the pressure is zero. For stationary loads these angles were assumed to be 0 and t; for a vari¬ able load they are time-dependent functions. The zero points can be obtained by writing = pjC/R)1 = u = ficj (1 — e cos 0)2 or u = 0. By writing out u from above, we have — 12e(l — G) sin 0 + 3«2(1 — G) sin 20 + 12E cos 6 — 3eE cos 26 = 0 With q = (1 — G)/E the above reads —4eq sin 6 + e2q sin 20 + 4 cos 6 — c cos 26 = 0 (7-75) An examination of Eq. (7-75) will show that in general for E = 0, 6 = it will always lie between the two zero points 0i and 02. For E > 0, the pressures are positive around 0 = 0 or hmin. For E < 0 the pressures are positive around 0 = t. Equation (7-75) can be rewritten as qt(4 sin 0 — e sin 20) — (4 cos 6 — e cos 20) = 0 By expressing the double argument in terms of single-argument functions, we have 2(qe sin 0 — cos 0)(2 — « cos 0) — e = 0 The simplification now introduced is that c cos 0 is small as compared to 2, and we have qe sin 0 — cos 0 — ^ = 0 By using the substitutions 2 tan 0/2 . 1 - tan2 0/2 sin 0 = -———' /rt cos 0 = ^- 1 + tan2 0/2 1 + tan2 0/2 we obtain the quadratic equation (i - j) tan* | + 2qt tan | - ^1 + = 0 where the roots are 4 0 _ — qt ± y/q2€2 + 1 — e2/16 tan 2 r^/4
Squeeze Film and Dynamic Loading 241 By neglecting c2/16 and rearranging the terms, we obtain for the two zero points 0i and 02 tan 9-f = ± 6 (l + hF ?] (7-76) which as seen are spaced tt radians apart. We are now to integrate the pressures from 0i to 02 for horizontal and vertical components along the coordinate axis of Fig. 7-13. These are given by Wv = — f pty) sin yj/LR d\fr J Qi Wx = — f p{yp) cos yf/LR dyj/ J 0i From Pit) = and t = 0 + y we also have p- - -1 wkp jttm ““ <* + ’)" - 5MW-F- (7’77“> p- - -1 wm Sn<,) c“ +7)" ’ s sot '"• (7-77t) where Fx and Fy represent the respective integrals or, with the sum of the angles expanded, are given by Fv — —/II(0)(sin 0 cos y + cos 0 sin 7) dd Fx = —/II(0)(cos 0 cos 7 — sin 0 sin 7) dd The evaluation of these integrals involves a good deal of mathematical manipulation, and for their details the reader is referred to Ref. 3. The result of these integrations is for E > 0 given by
242 Theory of Hydrodynamic Lubrication and for H < 0 Fy — cos 7 \ , x/ 6irt(l - G) ( k \ (2 + «2) (1 — .*)«V* + h) / \ Fx sin 7 sin 7 + (2 + t2)(l - «2)W(1 - (■) [4fc<2 “ (2 + <2)lr k + (7‘79) COS 7 where k = (1 - [(4^)’ + ^ By setting E = G = 0, the above equations can be reduced to the stand¬ ard Sommerfeld solution for a 0 — t positive pressure wave, or F = 6rfe F = 12«2 n ROx v (2 + €2)(1 - t2)^ 1 (2 + c2)(l - e2) W ; and the attitude angle is <t> = arctan -■ * jy* (7-81) As mentioned above, the analytically simpler technique of treating Eqs. (7-78) and (7-79) is to find the form of loading W = W(t) cor¬ responding to a given locus e = «(£) and y = 7(0, and our main interest in the following section will be concentrated on an analysis of sinusoidal and elliptical orbits. 7-14. Solutions for Prescribed Loci. 1. Circular Locus of Uniform Velocity. If the locus is circular, e = const, and if the velocity of travel of shaft center is uniform, 7 = cot and dy/dt = const = co. We thus have E=0 0=2— CO F» — cos cot _ (1 _ 2 —) G,rf ' Fx sin cot sin cot ~ | (l “ 2 T?) | - (7-82) (2 + €«)(1 - **)\ COS cot
Squeeze Film and Dynamic Loading 243 or for the resultant load capacity D V I O I 6€[7T2 — €2(tT2 — 4)P* (n _ . M \C) I" ^ (2 + <2)(1 - «2) ^ ^ Again, as in all previous analyses, the load capacity is zero when Wi = Hco. 2. Unidirectional Sinusoidal Locus. Let the equation of the locus be given by € = 6o H“ a sin wpt where the motion takes place along a straight line with y equal to either 0 or x. From these conditions we have 2 E = - aoip cos copt G = 0 a) With y equal to either zero or 7r, sin y = 0 and Eqs. (7-78) and (7-79) become for E > 0 fore / k + 3 \ (2 + ,*)(i - «*)>* yFTJi)008 7 ,7 3# l,,, k + 31 ( } (2 + «2)(1 - [ + ( + ‘ )lr k + % J cos 7 and for E < 0 Ei &7T6 / k \ Fw ~ (2 + .*)(1 - .*)» (fc + C0S 7 f--(2+.w-.■)*■ [4fa!-<2 + ■’>*thi]cos 1y (7-85) Two specific examples in which wp = w are given in Fig. 7-14. The diagrams represent the instantaneous resultant hydrodynamic forces and thus the external loading required to produce a sinusoidal motion of shaft center given by either e = 0.6 sin wt or e = 0.3 + 0.3 sin wt. It should be noted that the frequency of the marked orbit is that of journal rotation. The angular values marked on the diagram of the load vector when substituted in the expression for e will yield the corresponding position of shaft center. 3. Elliptical Orbits. The parametric equations of an ellipse are c sin y — ab sin wt € cos y = a cos wt
244 Theory of Hydrodynamic Lubrication where a and b are respectively the major and minor semiaxes and a is ± 1. Here too the load frequency is assumed to be that of shaft rotation. The ellipse, of course, is nothing more than a simultaneous sinusoidal motion in two directions, and we can distinguish two cases: those in which the orbit progresses either in the same direction as the load or in Fig. 7-14. Loading corresponding to sinusoidal journal locus: (a) force diagram for a journal locus e = 0.6 sin a>t; (b) force diagram for a journal locus « = 0.3 + 0.3 sin w/. (After H. H. Oil; by permission of G. Leemann Verlag, Zurich.) the opposite direction. By assigning to <r the value of either +1 or — 1, this can be taken into account in a single expression. From the expres¬ sions for c and y given above by noting that b 1 dy b o) tan y = a - tan ut ~ ~ a cos2 y dt a cos2 a>t we have L, /k2 sin 2u)t n 2ah t2 + 2oab h = (62 — a2) G = a — - q = €2 (b2 — a2)c sin 2wt These can now be used in Eqs. (7-78) and (7-79) to obtain the vertical and horizontal load components for the respective positions of the shaft center as given by the various values of u>t. The loads for the apexes A and B can be easily obtained from Eq. (7-82) by setting
245 Fig. 7-15. Loads corresponding to elliptical journal loci. (After H. H. Oil; by permission G. Leemann Verlag, Zurich.)
246 Theory of Hydrodynamic Lubrication For A: y = 0, e = a p _ for (a — 2ab) y (2 + a2)(l -a2)* For B: y = t/2, e = b Fv= - 12b\b — 2oa\ Fx = - 12a|(a - 2ab)\ (2 + a2)(l - a2) 6ir(6 — 2<ra) (2 + 52)(1 — 62) X x (2 + 52)(1 - 62)* For the points opposite to A and B the same expressions apply with their signs reversed. -1 170° 160° 150° 0 =o.4 u 90° 120° -I 170° 16(£150° Vffe 1121 120° 5 30° 60° 90° a=0.2 Fig. 7-16. Loads corresponding to elliptical orbits with b = a/2, <r = 1. (After H. H. Ott; 6j/ 'permission G. Leemann Verlag, Zurich.) Figure 7-15 gives a series of load diagrams for elliptical orbits with a fixed horizontal axis of a = 0.6 and a variable vertical axis ranging from 0 to 0.8. The following general comments can be made: The load diagram is symmetrical with respect to its 0-tt axis. The direction of rotation of the load is the same as that of the shaft orbit except in the range 0 < b < a/2, when the load rotates in the opposite direction. At b = a/2 the load diagram has a zero point and the direc¬ tion of the load is nearly along a straight line. To show that this is true not only in the case where a — 0.6, we have in Fig. 7-16 the shaft orbits for arbitrary values of a with b == a/2. It is seen that the direc¬ tional oscillations of the load are always small. Thus a load required to produce an elliptical orbit with the minor axis half that of the major
Squeeze Film and Dynamic Loading 247 axis is given by the diagram of Fig. 7-17. The diagram in the x-2x range will have an identical shape except for a reverse sign. When a = 6, as seen in Fig. 7-15, the load diagram becomes a circle. 7-15. Solutions for Prescribed Loads. For low eccentricities Eqs. (7-78) and (7-79) can be further sim¬ plified. In the range 0 < e < 0.5, the value of (2 -J- «2)(1 — c2)* varies between 2.00 and 1.95 and we can write (2 + €2)(1 - t2)* « 2.0 For small«it also results that k k + 3 k + H k + y2 1 Fig. 7-17. Loading as a function of $ for elliptical orbits with b =* a/2, o — 1. {After H. H. Ott; by 'permission G. Lee- mann Verlag, Zurich.) With these modifications the two cases E ^ 0 yield the following single set of equations \ — cos y = — 3ire(l — G) \ F, \ / sin y - 3 2e2 [<■- Gy + (‘i \/ k sin 7 (7-86) cos 7 We shall consider a unidirectional sinusoidal load given by Fy = aCi sin ut Fx = 0 Here again it should be noticed that the load frequency is the same as that of journal rotation. When this is used in Eq. (7-86), we have sin 7 3t«(1 - G) cos 7 - 3 |k yjd - or- + (^) + ttA’J s: - 3 [k yjd - GY + (£j + irE j cos 7 = 0 (7-87) 3tt€(1 — G) sin 7
248 Theory of Hydrodynamic Lubrication These two equations can be solved for G and E to give G = 1 — Fj,~S r E = - F‘ 7 ± (7-88) Sice ott 3tt2 where 4e2 was ignored against the value of it2. We shall now revert from the polar coordinates c, y to the cartesian system. By setting x = x/C and y = y/C, we have €2 = x2 -J- y2 tan y = j T = wt and from the definitions of G and E r _ 2 dy _ _0gdx_ ,0xdy U co dt e2 dT e2 dT 2 de_ xdx ydy h a>dt e dT € dT By replacing E and G by the values given in Eq. (7-88), we obtain 2* + 2y & = -yFo - \ (i* + yWo\ . . (7-89) -21/g + 2xg= (** + fl*) - where Fo = Fv/3t By solving Eqs. (7-89) for the first derivatives, we obtain -y-2x\Fo\ 2^ = x-^y\F,\-Fo (7-90) where Fo = <r(Ci/3ir) sin T = <rCo sin T. As a first approximation we neglect the terms containing Fo and, by writing x/3?r = x0, y/3t = y0, we have from Eq. (7-90) 2 g? = -yt 2 ^|° = x„ - „c. sin T (7-91) By differentiating the first equation and using it in the second, we obtain 4 + Xo = aCo sin T The solution of this equation is, by substitution of xo = B sin T} easily seen to be Xo = — livCo sin T and from (7-91) yo = %crC0 cos T (7-92) which are the equations of an ellipse with b = a/2.
Squeeze Film and Dynamic Loading 249 We now proceed to refine solution Eq. (7-92) by writing x = xo + xoi y = 2/0 + 2/oi where xoi and yoi are now the new unknowns. When x and y are placed in (7-90) with x0 and yo as given by Eq. (7-92), we obtain 2 w = ~m ~ \|Fol(Xo + Xoi) 2 w = Xo1 _ \|fo|(2/0 + yoi) By proceeding as above, i.e., ignoring the terms (2/x)FoXoi and (2/t)Fo2/oi, we obtain for approximate solutions denoted by xx and yi 2 S? = ~Vl ~ \|Fo|xo 2 = Xl “ \ |Fo|2/0 (7'93) This set of linear nonhomogeneous equations does not differ essentially from Eq. (7-91) and can be easily solved. We can now proceed with a further refinement by writing Xoi = Xl + Xn 2/01 = 2/1 + 2/11 and these are now inserted in the set preceding Eq. (7-93) to yield 2 = — 2/11 — ~ |Fo|(xi + Xu) 2 = Xu — - |Fo|(2/1 H~ 2/11) This is a set of equations analogous to Eq. (7-93). We continue by writing £11 = x2 + x2i 2/11 = 2/2 + 2/21 and, ignoring the terms (2/x)FoXn and (2/x)F02/n, we obtain S 2 ' (7-94) 2A' = x*-l^' which is a set identical with Eq. (7-93). We can continue in that manner for any arbitrary number of steps, and the set of equations to solve will always be of this form: o 2 1»•» 1 2 ~77fi ~ -Ui ko|xt_i f , ’ (») The solution of any of these sets of equations is accomplished by differ¬ entiating the first equation with respect to T} solving for dyi/dT, and using
250 Theory of Hydrodynamic Lubrication the solution in the second equation. The resulting single equation is then of the form where F0 has been replaced by its proper value. Equation (7-96) is essentially the differential equation for forced vibra¬ tion with no damping in the system. Since the solutions will always be in the form of the particular solution k sin T and since x0 and yo have a period of 2t, all subsequent solutions will likewise have a period of 2*. The right-hand side of Eq. (7-96) can be written in terms of a Fourier series, so that The expressions above can be shown to be rapidly converging series, and accurate results are obtained if no more than two indices are taken, i = 0 and i = l. For i = 0, x0 and yo are given by Eq. (7-92). For i = 1 we have, by considering the right-hand side of Eq. (7-96) with x0 and y0 replaced by their proper values, the following: d2r- 2 4 d + = t Co|sin T|^-1 — 7r 0 <TT(|sin T\Xi-l) (7‘96) 4 -T^1 + *» = / Ain cos nT + Bin sin nT n «■ 1 and thus the solution to this equation must be of the form tl From Eqs. (7-97), then, we obtain
Squeeze Film and Dynamic Loading 251 The total solution then is oo x = xo + xi = —licCo sin T + ^ <rC02 ^ cos nT 1,3,5 (7-98) 2/ ^ T , IB ,2V 2n2 + 1 y = l/o + yi = /3<tCo COS 7 + u ^ 2 3n(n* - 4)(4n2 - 1) sin nT 1.3.5 The expressions above are rapidly converging series, and good accuracy is obtained if only three terms of the series are retained. With these approximations the equations above become X = -HffC'o sin T - ^,<rCV cos T + -^<rC02 cos 3T 97T2 1 / 07r2 y = %<rC„ cos T — <rCo2 sin 1' + <rCV sin 371 97T2 1,0/07T2 (7-99) The first terms on the right-hand side of Eqs. (7-99) can be seen to repre¬ sent the equations of an ellipse. The remaining terms represent the Fig. 7-18. Accuracy of solution (Eqs. (7-99)). (After H. H. Ott; by permission G. Lee- mann Verlag, Zurich.) extent of “distortion” of this ellipse. They are sufficiently small, at least for small e, to retain for the locus of a shaft subject to a sinusoidal load the essential features of an elliptical orbit. A check can be made on the extent of validity of solution (7-99) by putting the orbit as given by x and y into Eqs. (7-78) and (7-79) and recalculating the resultant Fx and Fy. For complete accuracy we must have Fx = 0 and Fy = Ci sin cot. Figure 7-18 shows the extent of deviation due to the assumption made at the beginning of this paragraph. Solution (7-99) seems to be restricted to Ci < 1.5, which corresponds to an eccentricity no larger than 0.1.
252 Theory of Hydrodynamic Lubrication To improve upon this rather narrow restriction, we shall return to Eqs. (7-78) and (7-79), which carry no restriction on the magnitude of e, but we shall retain the simplification (2 -f- c2)(l — €2) « 2.00. This is valid for eccentricities up to 0.5. The resulting two equations are then for E > 0. ri cos y K y \ T. . / \ - /fc + 3\' ' 3£ y+ h) 2d -.») sin 7 sin 7 l 4 kS + (2 + ««)x (7-100) A: + 3 + y2j \ COS 7 and for E < 0 „ - cos 7 “ / )- -3"(I - (iT«) F‘ \in 7 sin 7 + 2(r?lii[4h,-(2 + <,»'iTTs] <7-101> COS 7 We shall assume a solution of the form x = —\ioCo sin T + k&Co2 cos T + k2oCo2 cos 3T . y = 2/z<jcq cos T + kzoCo2 sin T + hoCo2 sin 3T U-UW; where A;i, k2} fa, and fa are to be determined from the conditions that Fx = 0 and Fv = oC\ sin T. These solutions, owing to the complexity of Eqs. (7-101) and (7-102), were obtained by numerical means with the calculations carried out at four locations: T = 0°, 45°, 90°, and 135°. With the constants evaluated, solutions (7-102) read x = -^oCo sin T - 0.2868(rCo2 cos T + 0.0195aC02 cos 3T y = 2\£ac0 cos T - 0.3920(rCo2 sin T + 0.0861<rCV sin 3T U Table 7-3 gives the values of Fx and Fy as obtained from Eqs. (7-100) and (7-101), and Fig. 7-19 gives an evaluation of the accuracy involved. It seems that the use of Eqs. (7-100) and (7-101) extended the validity of our solutions to Ci < 3. Figure 7-20 gives the resulting locus of shaft
Fig. 7-19. Accuracy of solution [Eqs. (7-103)]. (After H. H. Ott; by permission G. Lee- mann Verlag, Zurich.) Fig. 7-20. Journal locus under sinusoidal loading. (After //. H. Ott; by permission G. Leemann Verlag, Zurich.) 253
254 Theory of Hydrodynamic Lubrication center, and we see that for C\ = 3 the maximum allowable eccentricity is about 0.2. Figure 7-20 also shows part of the locus of shaft center for steady loads whose magnitudes are equal to the amplitudes of the sinus¬ oidal loads. The eccentricities for the steady loads are in all cases higher than for the sinusoidal loads, which indicates again the higher load capacity of dynamically loaded bearings. Table 7-3. Load Components for Sinusoidal Load [Eqs. (7-103)1 cx copt = 0 <opt = 45° (opt = H cO o 0 (opt = 1 = 135° Fy Fx Fy Fx Fy Fx Fy Fx 1 0 0 0.710 0.002 1.001 0.001 0.696 -0.003 2 0 0 1.418 0.005 2.070 0.006 1.350 -0.019 3 0 0 2.156 0.043 3.270 0.015 2.007 -0.056 4 0 0 . 2.917 0.092 4.670 0.020 2.519 -0.104 SOLUTIONS FOR FINITE JOURNAL BEARINGS 7-16. Dynamic Loading. The two-dimensional equation for dynamic loading is r (~ Pi + 7r(-f) = (iUT+ 12Ko dx \n dx) dy\y. dx) dx The process of converting this equation into dimensionless form follows the steps in Chap. 4 with V0 supplying a new set of input parameters. By using the substitutions . * , * ch p(C/Ry *=/e z = Z72 h c n-—^~ and remembering that V0 = dh/dt = C(dh/dt), we obtain where the angle ^ is as given in Fig. 7-13. Now by the use of the familiar transformation u = h* n Eq. (7-104) transforms into dH ,(DY &u _ J_ jl [MV , (D\* /dh\] dr + \l) dV 2h2 \2 my \L/ \d~z) J
Squeeze Film and Dynamic Loading 255 If the angular coordinate is rewritten as 0 = (\j/ — y) and the expression for h = 1 — € cos 0 is written out, the basic differential equation becomes S+(?)2 S+a(e’°u - b{9'i) (7*io5) with ^ r€2(sin2 0 — 2 cos2 0) + 2e cos 0] 4(1 — € cos 0)2 H6,t) = (6C0S t« sin 8(1 -G) - E cos 9] where E = - ^ and G = - ^ . (a at a) at Although the resulting differential equation u = is a function of three variables, the coordinate of time appears only in connection with the parameters E and (7, and thus Eq. (7-105) can be solved only for 0 and z. When the path is known, c and y and their derivatives are known, and the pressures arid resultant load can be calculated. This, as pointed out before, is the easier approach, but one that is the reverse of practical situations when the loading is known and the locus of shaft center must be obtained from the differential equation. The boundary conditions are p(^) = p(*,-t;) = 0 p(8hz) = 0 »<»•■« - -0 where 0i and 02 are the angles corresponding to the beginning and end of the pressure wave. However, such boundary conditions present formi¬ dable difficulties, and the simplification is made that the pressure profile starts and ends at 0 and 7r, respectively. The four boundary conditions thus are p(0,L/2) = p(0,-L/2) = p(0,2) = p(ir,z) = 0 In terms of the variable u these boundary conditions become w(0,l) = u(0f — 1) = u(0,z) = u(tt,z) = 0 Differential equation (7-105) with all its terms written out reads d2u 602 + (r)2 S - 4(i - f cos ey [t’(sin2 e - 2 cos2 e) + 2< cos e]u = J 6 -rru r€ sin 0^1-- ^ cos (7_106) (1 — € COS 0) J L \ 03 dt) 03 dt J
256 Theory of Hydrodynamic Lubrication This is a linear differential equation and thus the results of the different terms of the right-hand side of Eq. (7-106) can be superimposed. The right-hand side consists of three terms: 6h~^e sin 0, which is the contribution of the journal rotating about its center with an angular velocity w 2 dy — 6h~V- ~jt sin 0, which is the term due to the tangential velocity of co dt journal center — 6/H* - ^ cos 0, which is the term due to the radial velocity of journal o) dt center Upon combining the terms in sin 0, we are left with two principal terms: 6fi-#(l - (?) sin 0 = 6h-»G' sin 0 (?' = (1 - (?) and —6hr^E cos 0 The general solutions can now be obtained by obtaining results for E and (?' separately and then superimposing these special solutions. The force components in dimensionless form are Fx = J* Jj II cos 0 dd dz Fy = J* Jj II sin 0 dd dz with the range of integration extending only over p > 0. The resultant load capacity and the phase angle <£, that is, the angle between the line of centers and the resultant instantaneous load, is So K?'| where Fx = — ( li - V(iSi)'+(iSi)’ <7-107> fJLO) \R The sign of (?' is determined by the value of <f> as follows: 0 < <t> <TT (?' < 0 7T < <f> < 2tt G' > 0 (7-108) <f> = TVK G' — 0 The above equations were solved for the case of L/D = H by means of mathematical relaxation. The results relating So to the eccentricity and phase angle with q = E/(G') as a parameter are given in Figs. 7-21 and 7-22.
radians Squeeze Film and Dynamic Loading 257 Fig. 7-21. Magnitude of So/IG'I = /(«,#,G). Numbers refer to E/\G'\ ratios. Fig. 7-22. Phase angle <j> = /(«,/£,(/). Numbers refer to E/\G'\ ratios.
258 Theory of Hydrodynamic Lubrication As mentioned previously, the calculation of the resultant load when the journal path is known is relatively simple; the necessary data can be obtained directly from Figs. 7-21 and 7-22. For a constant load, for example, q = E/G' = 0 and the line 5 = 0 can be used to obtain So and <t> for a given eccentricity. When the journal is rotating at a con¬ stant e with some angular velocity coL = dy/dt, q = 0 but the value of Sq/G' will depend on the value 1 — 2wl/w, which is seen to bear the same qualitative relation to load capacity as obtained in the one-dimensional Fig. 7-23. Loads corresponding to given position of journal center and value So. solutions. For a general orbit, e, y, de/dt and dy/dt have to be known. Figure 7-23 shows the relation of the load So to an elliptical path of journal center as given by the previously treated special kind of motion c sin y = 6 sin cot e cos y = a cos cot These are seen to be similar to the one-dimensional solutions of Fig. 7-15. The calculation of the journal path when the loading is known is more difficult. The relationship of path and loading is given by \G'\ fl VD’ c’7’ dt’ dt) * f2 V/)’ ^ dt’ dt) These relationships are given in Figs. 7-21 and 7-22. Analytically they represent a system of two linear differential equations in e and y.
Squeeze Film and Dynamic Loading 259 For an arbitrary load diagram the boundary conditions of thi system are periodic relations of e and y which must be satisfied by = + T) with similar requirements on ef y, de/dt> and dy/dt. The boundary conditions thus are de (0) _ de (T) <(0) = e(T) 7(0) = y(T) dt dt dy (0) _fdy (T) dt dt The problem has to be treated as an initial value problem. A starting e and <t> are arbitrarily selected, and the path which at the end of the cycle x4>wf--3W4 , P^W> COS ut Fig. 7-24. Journal locus under sinusoidal loading. must fulfill the above boundary conditions is calculated step by step. Since €(0) and 7(0) are assumed and <S0 and <t> are known from the load diagram, these values can be used to find the value of q = E/(G'). The sign of G' is determined from the value of <f>, as given before. Thus q = E/G' and S'0 = So/G' provide the values of E and G and conse¬ quently of de/dt and dy/dt.
260 Theory of Hydrodynamic Lubrication A sample solution for a unidirectional sinusoidal load given by the equation So = Sm cos mt is shown in Fig. 7-24 for a range of values of Sm. If we take the more general case with the frequency of load oscillation different from that of shaft rotation, that is, So = Sm cos out, we have Fig. 7-25, in which the curves are plotted for various values of o>l/« with a fixed Sm = 1. The relation between the maximum eccentricity and frequency ratio wl/g) is given in Fig. 7-26. o 1 >(jjpt = 0 *2 tu)ptz tt/4 +3 luipi- ir/Z x 4 Ivpt- 3tt/4 Jo * Sm MS "<•' = — COS Wft Fig. 7-25. Journal locus under sinusoidal loading of arbitrary frequency. Numbers indicate «/>/w ratios. Additional solutions for finite bearings in which the shaft center has radial and tangential motion are discussed in Chap. 8 in connection with the calculation of spring and damping functions of fluid films. Table 8-1 gives the eccentricity, phase angle, and load components as functions of the radial and tangential velocity components of Fo for various L/D ratios of the bearing. 7-17. Squeeze Films. For the case of squeeze films, U = 0 and the differential equation of the preceding section reduces to !(*•!)+!(“ 2)
Squeeze FUrn and Dynamic Loading 201 40e -40‘ -60° \ Uit - 0 A ^— C • 4 A V \l Sr 0.5 pS*=0.3 1 \ ^•0.1 0.25 0.5 0.75 1.0 1.25 Uip/di Fig. 7-26. Maximum eccentricity and phase angle for sinusoidal load¬ ing. So = 1, du>p/d((U) = const; — So = 1, « = const. Fig. 7-27. Load capacity of 180° journal bearing under squeeze films. Eccentricity rahc-* Fig. 7-28. Side leakage of 180° journal bearing under squeeze films. Fig. 7-29. Journal eccentricity under sinusoidal loading.
262 Theory of Hydrodynamic Lubrication We shall consider the case of a 180° journal bearing with the journal traveling along the line of centers. By considering both positive and negative eccentricities, the solution for the 180° bearing can be made applicable to a full bearing, since the latter is made up of two 180° arcs, Fig. 7-30. Maximum eccentricity under Fig. 7-31. Power loss under sinusoidal sinusoidal loading. loading. one arc subject to a positive and the other to a negative journal dis¬ placement. Under these conditions the above equation becomes de where e = Equation (7-109) is worked out6 by assuming a solution of the form 7rz sin n -j- cos md and evaluating the coefficients on a digital computer. Figures 7-27 and 7-28 give the resultant hydrodynamic forces and side leakage as a function of journal velocity and eccentricity. Figures 7-29 to 7-33 give relations for a sinusoidal mode of loading W(o)Pl) = Wo sin wpt The flows for the sinusoidal loading are to be understood as time averages for the complete cycle. The total flow Qin includes both the side leak-
Squeeze Film and Dynamic Loading 263 age and the flow out the axial ends of the 180° arc due to the pressure gradients. The power loss H too is the average for the cycle, and in the <0 I Cfc; tu CVJ I 0.1 0.01 0.001 - = sin T - Z/Z? = 0.1 / 0.2/ 0.4/ L 1 1—1— 0.6 y 1.0X i i i i 0.1 & 10 20 fiutp / R \2 2VP0 Fiq. 7-32. Average side flow under sinus¬ oidal loading. 3 PO <o s 0.1 * 0.01 ' Z/ZM.O W[T)-- IV0 sin T 0.6 0.4 : 0.2 ‘ 0.1 - 1 1 L_l_ —i—i i_i_ . i__ _. i ii 0.01 0.1 10 H-up /ft_\2/ L ZttPACiVD Fig. 7-33. Average total flow under sinusoidal loading. absence of journal rotation is due to the oscillation of the journal in the oil film as given by where T = apt. SOURCES 1. Archibald, F.: Load Capacity and Time Relations for Squeeze Films, Trans. ASME, vol. 78, pp. 24-35, January, 1956. 2. Burwell, J. T.: The Calculated Performance of Dynamically Loaded Sleeve Bearings, J. Appl. Mechanics, vol. 69, pp. A231-A245, 1947; vol. 71, pp. 358-360,1949. 3. Ott, H. H.: “ Zylindrische Gleitlager bei instationarer Belastung,” Verlag A. G. Leemann, Zurich, 1948. 4. Hahn, H. W.: Dynamically Loaded Journal Bearings of Finite Length, Conf. on Lubrication and Wear, Paper 55, London, 1957. 5. Swift, H. W.: Fluctuating Loads in Sleeve Bearings, J. Inst. Civil Engrs. {Lon¬ don), vol. 5, p. 161, 1937. 6. Hays, D. F.: “Squeeze Films: A Finite Journal Bearing with a Fluctuating Load,” General Motors Corp., 1960.
CHAPTER 8 HYDRODYNAMIC INSTABILITY 8-1. The Mechanics of Hydrodynamic Instability. Hydrodynamic instability is caused by forces generated in the fluid film of the bearing, so directed with respect to the shaft displacement as to propel the shaft in its whirling motion. To visualize better the forces which cause this whirling motion, consider Fig. 8-1, which shows the equilibrium attitude of a shaft rotating in a bearing and supporting a load W. Because this is the position of equilibrium, the resultant force of the fluid film on the shaft is the force F, which must be equal and opposite to W. The important thing to notice in Fig. 8-1 is that the force F is not in the Fig. 8-1. Rotating shaft in equilibrium Fig. 8-2. Unloaded shaft displaced from position. the bearing center. direction of centers O'O, but, rather, is at an angle <f> to the direction of centers. Now let us consider a rotating shaft which carries no load (W = 0) and suppose that it is momentarily displaced from 0 to O', as shown in Fig. 8-2. The film under these conditions exerts a resultant force F on the shaft, just as it did in Fig. 8-1. However, in this case there is no opposing force W, so the force F must be spent in accelerating the shaft and over¬ coming the frictional drag of its resulting motion. The movement of the shaft center O' in response to the force F obviously will not be toward the bearing center 0. Rather, the shaft center will be forced to move in an orbit around the bearing center, and as long as the centers do not coincide, some force F will be generated by the rotation of the shaft and the whirl 264
Hydrodynamic Instability 265 will continue. Whether the whirling motion becomes more pronounced, continues at the same amplitude, or dies out depends upon the angle <f> and the damping characteristics of the bearing and shaft system. The whirl frequency is set by the speed at which the shaft can pump the fluid around in the clearance and maintain the pressure pattern which produces the driving force F. Assuming laminar conditions and neglect¬ ing flow due to any pressure gradients (which in a lightly loaded bearing are very small), the average lubricant velocity is half the peripheral speed and in the direction of shaft rotation. Hence, the greatest fre¬ quency at which a pressure pattern can progress around the bearing is half the shaft speed. This is the speed and direction at which the shaft vibration occurs. We shall refer to this kind of instability as half¬ frequency whirl. Finally, consider the case illustrated in Fig. 8-3, in which the shaft supports a load W in the equilibrium position O'. Suppose the shaft, by some external shock, is momen¬ tarily given a secondary displacement to a new position 0". The fluid force F corresponding to this new position of the shaft is no longer equal and opposite to W. The vector difference between the forces F and IF is a force F2 which can cause the shaft center to whirl (in the direction of shaft rotation) around its equilibrium position O' at a speed nearly equal to half the shaft rotational speed—just as in the case of the unloaded shaft discussed above. As before, the per¬ sistence of such a whirling motion will depend upon the damping characteristics of the system. Another phenomenon, system resonance, can join with half-frequency whirl to produce a vigorous vibration. When the rotational speed is about twice the actual system first critical, the system will build up in resonance at a frequency equal to the system first critical frequency. This form of resonance, referred to as resonant whip, may be defined as a resonant vibration of a journal in a fluid-film bearing which, for low eccentricity ratios, sets in at approximately twice the actual first system critical and persists at higher speeds with the frequency of vibration approximately equal to the first system critical regardless of running speed. Here too the motion of the shaft center is in the same direction as shaft rotation. Resonant whip is a self-supported vibration, as is half-frequency whirl. In the case of resonant whip, the vibration is supported by the fluid-film action, while the frequency is controlled by the system critical speed. 8-2. Hydrodynamic Forces on Journal. The general form of the Reynolds equation for incompressible fluids may be written Fig. 8-3. Loaded shaft displaced from its equilib¬ rium position.
266 Theory of Hydrodynamic Lubrication d_ dd £(1 -1- € cos 6) ] + ft’A[(i + ,cose),xg] [u> — 2a) sin 6 -f- 2e cos 6] (8-1) Here the journal speed is a>, and the velocity components of the shaft cen¬ ter in the radial and tangential directions are Vr (= Ci) and Vt (= Cea). As explained in Chaps. 3 and 4, Eq. (8-1) is subject to the following boundary conditions: Solution of Eq. (8-1) yields the pressure distribution in the bearing side B, one obtains the resultant force which the fluid film exerts on the journal, and thus one may evaluate Fr,Fh the components of the force in the radial and in the tangential directions. These components can be expressed in the following form: The negative sign in the first equation is due to the fact that Fr is measured positive in the direction of increasing e, while (at least for Vt — 0) the dimensionless radial force actually turns out to point in the direction opposing the radial displacement. The right-hand side of Eq. (8-1) is linear in w, a, and i and can be put in the alternative form It would appear that the linearity of the Reynolds equation in U, Vr, Vt should lead to the conclusion that the equation can be integrated separately for a stationary center (with Vr = Vt = 0), then for a radial velocity Vr of the center (with U = 0, Vt = 0), finally for a tangential velocity Vt (with U = 0, Vr = 0), and the pressures (and hence also the forces Fri Ft) superposed to obtain the joint effect of Uf Vr, Vt. These conclusions, however, are vitiated by the nonlinear requirement that p = 0 over the area where cavitation occurs Sp anc* aa = 0 on boundary B of the cavitation area outside of B (p = 0 inside of B). By integration of the pressures out- (8-2) 6*i — (eCU — 2RVt) sin 6 + 2RVr cos 6]
Hydrodynamic Instability 267 p > 0 and the fact that the cavitation curve B will, in general, change in a manner depending on all three velocities, U, Vr, and Vt. Nevertheless, an examination of the right-hand side of the Reynolds equation shows that at least U and Vt enter in the form (eCU - 2RVt) sin 6 or in the equivalent form (w — 2d) sin 0. If Eq. (8-1) is integrated for a given L/D and the forces Fr, Ft in Eqs. (8-2) are evaluated for Vt = 0, the equations reduce to the following form: Fr = -\ufrM Ft = X0*f|(€,O (8-3) where /r, /< are dimensionless measures of the radial and tangential forces and are functions of the eccentricity ratio e and e'. The latter is a dimen¬ sionless measure of the rate of change of eccentricity: ' = Jl = 1 yr * o> dt it 2CN while Xu has dimensions of force, X being defined by . _ uLR (R\ _ SF A — IT For nonzero Vt or d the forces may be derived from the solutions for the case where Vt = 0. It is only necessary to replace U in Eqs. (8-2) by U - 2~ = Rw - 2a Ce or to replace w by w(l — 2a'), where a! = d/w = RVt/tCU. This requires also that the dimensionless parameter e' be divided by (1 — 2a'). Hence, for Vt ^ 0, the fluid film forces become Fr = -X«(l - 2a')/r(€,«') , . Ft = X«(l - 2a')/t(e,e') ^ where I' = ,—(8-5) 1 — la Numerical integrations of Eq. (8-1) have been carried out for plain journal bearings. The resulting dimensionless forces fT and ft of Eq. (8-3) are given in Tables 1 and 2 as functions of e, e' and L/D. The results extend the available solutions of statically loaded plain journal bearings given in Chap. 4 to the dynamic cases by including the additional param¬ eter c' in the analysis.
Table 8-1. Incompressible Plain Journal Bearing, L/D = 1 (No tangential shaft center velocity) Nomencloture -fr e a> -K, +ft S' u O' ~lf , O .E F + ^ ^ 3 z3(- < re' fr ft / 180 — <t> 0.1 2.0 10.0 0.090 10.0 179.5 1.5 7.30 0.089 7.30 179.3 1.0 4.80 0.088 4.80 179.0 0.5 2.25 0.087 2.25 177.8 0 0.001 0.075 0.075 91.0 -1.0 - 5.00 0.067 5.00 0.8 -2.0 -10.00 0.065 10.00 0.4 0.3 2.0 20.0 3.10 1 20.2 171.2 1.5 15.0 3.08 j 15.3 168.4 1.0 10.0 3.00 1 10.5 163.3 0.5 5.00 2.96 i 5.81 149.4 0 0.798 2.43 ! 2.56 108.2 -1.0 - 2.50 1.75 3.05 35.0 -2.0 - 5.50 1.69 5.75 17.1 0.5 2.0 40.00 7.30 40.7 169.7 1.5 29.50 6.80 30.3 167.0 1.0 19.50 6.45 20.5 161.7 0.5 10.50 5.70 12.0 151.5 0 2.90 4.77 5.58 121.3 -1.0 - 1.55 2.66 1 3.08 59.8 -2.0 - 3.90 2.48 4.63 32.4 0.7 2.0 102 15.0 103.3 171.6 1.5 80.0 14.5 81.4 169.7 1.0 52.0 13.7 53.9 165.2 0.5 27.0 12.0 30.1 156.0 0 9.67 9.08 13.3 136.8 -0.5 1.60 5.75 5.97 105.5 -1.0 - 1.00 3.80 3.93 75.3 -1.5 - 1.85 3.24 3.73 60.3 -2.0 - 2.98 3.06 | 3.33 45.7 0.05 2.0 2,000 70.0 2,000 178.0 1.5 1,495 69.0 1,496 177.4 1.0 1,000 67.0 1,002 176.2 0.5 545 58.0 548 173.9 0 112 38.0 118 161.3 -0.5 13.0 13.5 19.4 133.9 -1.0 3.5 6.50 ! 7.38 118.3 208
Hydrodynamic Instability 2(59 Table 8-2. Incompressible Plain Journal Bearing (Equal radial and tangential shaft center velocities) Tta L/D - 0.5 L/D - 1.0 L/D - 1.5 fr ft / 180-0 fr ft / 180-0 fr ft i / 180-0 6.0 11.2 0.2 14.5 140.4 33.2 31.5 45.0 136.8 64.0 56.5 86.0 138.0 4.0 7.4 6.4 0.8 130.1 23.0 21.4 32.1 138.2 43.0 38.0 57.4 138.5 2.0 3.6 3.4 4.0 137.0 11.7 10.0 16.0 137.0 20.0 20.5 28.6 134.3 1.0 1.4 1.6 2.2 130.3 4.5 4.0 6.7 132.6 8.4 10.8 13.7 127.0 0 0 0.2 0.2 08.3 0.1 0.7 0.7 100.1 0.3 1.3 1.3 102.4 6.0 20.3 13.0 24.1 147.4 62.0 40.0 73.8 147.2 104.0 72.0 126.5 145.3 4.0 13.6 0.0 16.2 146.3 41.2 27.5 40.5 146.3 68.5 48.5 83.0 144.7 2.0 7.2 5.0 8.8 145.4 10.5 15.0 24.6 142.4 34.0 26.0 42.8 142.6 1.0 3.6 2.8 4.6 142.0 7.8 7.8 11.0 135.2 17.5 15.3 23.2 138.8 0 0.3 0.8 0.8 111.8 0.8 2.4 2.5 108.2 1.6 4.0 4.3 112.1 - 0.6 -0.3 - 0.1 0.3 342.3 - 0.6 - 0.4 0.8 330.5 - 1.0 - 0.5 1.1 331.5 - 2.0 -1.0 - 1.7 2.0 300.8 - 2.0 - 4.3 4.8 205.5 - 3.2 - 8.5 0.1 200.6 - 4.0 -2.0 - 4.2 4.6 205.5 - 4.4 -12.8 13.5 288.8 - 6.0 - 21.0 21.8 286.0 -18.0 -8.7 -20.5 22.3 203.0 -10.6 -65.0 67.0 286.8 -26.0 -105.0 108.2 283.0 6.0 45.0 10.6 40.1 156.5 123.0 56.0 135.3 155.5 182.0 01.5 203.7 153.3 4.0 30.0 13.8 33.0 155.3 80.0 30.5 80.2 153.7 122.0 63.5 137.5 152.5 2.0 15.6 7.8 17.4 153.6 30.0 21.8 44.7 150.8 62.5 35.0 71.6 150.7 1.0 8.1 4.7 0.4 140.0 17.3 11.0 21.0 145.5 31.8 22.0 38.7 145.3 0 1.1 1.6 1.0 125.6 2.0 4.8 5.6 121.3 4.8 7.4 8.8 122.8 - 0.6 -0.2 0.2 0.3 35.6 - 0.6 0.6 0.8 47.5 - 0.7 0.8 48.3 48.0 - 2.0 -0.8 - 1.2 1.4 303.2 - 1.5 - 2.3 2.7 303.0 - 2.2 - 6.0 6.4 280.7 - 4.0 -1.4 - 2.2 2.6 301.2 - 2.8 -10.6 11.0 284.8 - 3.6 - 17.0 17.4 282.0 -18.0 -4.8 -10.2 10.8 284.0 -10.3 -58.5 50.4 280.0 -13.7 - 05.0 06.0 278.2 6.0 136.0 36.5 140.7 165.0 200.0 87.0 302.8 163.3 420.0 125.0 438.2 163.4 4.0 02.5 26.0 06.1 164.3 205.0 62.5 214.2 163.0 278.0 00.0 202.2 162.1 2.0 45.0 15.0 47.4 161.6 08.0 36.5 104.4 150.6 130.0 53.5 140.0 150.0 1.0 10.5 0.6 21.7 153.8 42.5 21.0 47.4 153.7 67.5 32.0 74.7 154.6 0 4.1 3.5 5.4 130.7 0.7 0.1 13.3 136.8 13.1 13.0 18.5 135.5 6.0 332.0 58.0 337.6 170.1 650.0 124.0 661.8 160.2 780.0 165.0 707.5 168.0 4.0 220.0 42.0 224.0 160.2 425.0 88.0 433.6 168.3 520.0 118.0 532.0 167.2 2.0 113.0 25.0 115.8 167.5 215.0 52.5 221.4 166.3 265.0 71.0 274.4 165.0 1.0 56.5 16.0 58.7 163.5 112.0 34.3 116.6 163.0 133.0 46.5 140.7 160.8 0 8.8 5.8 10.5 146.5 18.3 13.1 22.5 144.4 23.6 18.0 20.7 142.7 - 0.6 0.1 0.0 0.0 06.3 0.2 2.4 2.4 03.0 0.3 3.5 3.5 04.2 - 2.0 -0.5 - 0.6 0.7 300.2 - 1.1 - 1.3 1.7 300.6 - 1.5 - 2.5 2.0 301.0 - 4.0 -0.8 - 2.4 2.6 288.7 - 2.0 - 7.2 7.5 285.4 - 2.5 - 11.2 11.5 282.5 -18.0 -0.6 -18.6 18.6 271.8 - 0.4 -54.5 54.5 270.5 - 0.1 - 84.0 84.0 270.0 8-3. Threshold for Half-frequency Whirl. By employing the hydro- dynamic forces generated within the fluid film, it has been shown8 4* that the threshold of instability may be represented by M“'(i + /k)<4 <8J” where 1/K2 is the radial fluid-film resilience and l/k is the rotor resilience. * Such superscript figures indicate references listed under Sources at the end of the chapter.
270 Theory of Hydrodynamic Lubrication It can be shown that this inequality applies to bearings which possess a high degree of symmetry and that the results apply to both compress¬ ible and incompressible fluids. Neglecting friction forces, it has been shown in Chap. 7 that the result¬ ant film force due to journal rotation in a plain bearing is at right angles In the limiting condition no pressure can exist in the film, hence side leakage is zero, and it follows that The same result is obtained regardless of where points A and B are taken. 8-4. Forced Vibration of Vertical Rotor. Let us consider a vertical rotor supported by two plain journal bearings. The rotor consists of a flywheel fitted to a shaft and, to simplify the analytical work, it will be assumed that the rotor mass is concentrated at a point situated halfway between the bearings and, further, that the bearings are identical. The center of gravity of the rotor is offset a distance S from the center of the shaft. For the amplitudes of vibration of the center line of the shaft at the bottom bearing, flywheel, and top bearing we will use the symbols eiy e2, and 63, respectively. The actual bending deflections of the shaft will be noted by y. The relation between the displacements at various points on the shaft will be as indicated by the displacement diagram of Fig. 8-5. The amplitude of whirl at the flywheel is partly due to the displacement of the shaft in the bearings and partly due to the deflection of the shaft. Further, it will be seen that the amplitude of vibration of the center of gravity of the rotor is the vector sum of e2 and 8. to the eccentricity and is just balanced by the resultant force due to whirling when the whirling frequency is one-half rotational frequency. This result represents the upper limit of whirling frequency for an unloaded Fig. 8-4. Whirling journal. B ideal bearing. The same result can be deduced more directly by considering the continuity condition that the volume of lubricant passing through the film at some point A} Fig. 8-4, must equal the volume passing point B plus the volume required to fill the void of the receding journal. Hence a 2 (8-7)
Hydrodynamic Instability 271 Figure 8-6 shows the force diagram for the rotor. Let us denote the angular velocity of the shaft by w, and the centrifugal force acting on the rotor will be given by the following expression: CF = Mu}2(e2 + 8) (8-8) Since there is no damping in the shaft, the phase angle of the shaft deflection must be the same as that of the applied centrifugal force, and the vectors representing these two quantities must be parallel. As shown by the force diagram, the centrifugal force is composed of the two com¬ ponents Mo)2B and Mw2e2 parallel to the amplitudes 8 and e2, respectively. Since the centrifugal force acting on the rotor is equally divided between the bottom and top bearings, the force acting on each of these Mdi18 Md2(£2+8)\ Ic 0 Fig. 8-6. Forces acting on rotor. bearings will be equal to CF/2. Also, with the bearings assumed to be identical, the shaft displacements must be the same in both bearings. We will therefore drop the suffixes and write e\ = e2 = e. Since the forces applied to the bearings are of constant magnitude and rotate with a constant angular velocity, the relation between force and displacement can be determined from the constant load characteristics of the bearings, for it has been shown previously that introducing a value for U makes no difference to the principle of solution. Consequently, the use of curves obtained from the constant load calculations can be extended to cover the case of a constant rotating load. If we consider the disturbing force as the unbalance force, then it rotates with the same angular velocity as the journal, that is, a = w. If use is made of the load number / (Table 8-1) and the rotation is taken into account, the force CF will be given by the following expression: ? = i^LD©2/ei<“,+*> (8'9) For convenience, the attitude angle <f> has been chosen to be positive when the force on the fluid film leads the displacement, i.e., for negative Fig. 8-5. Displacement diagram for rotor- bearing system.
272 Theory of Hydrodynamic Lubrication load numbers. Here, / and 0 are a function of the eccentricity ratio e as given in Table 8-1. Let us introduce the symbol @ for the phase angle difference between vectors 8 and e2 -f 8 as shown in Fig. 8-5. By resolving the displacement vectors in the x and y direction, we arrive at the following two equations: e sin 0 = 8 sin P (8-10) 8 + e2 = e cos </> -f 8 cos p + y (8-11) The deflection of the shaft is obtained from simple beam theory and is given by _CFZ3 y 48 EI ( ^ If use is made of Eqs. (8-8), (8-9), and (8-12), Eq. (8-11) can now be written as «.♦ + ««»« (8-13) The unknown p can be eliminated between Eqs. (8-10) and (8-13), and as a result we obtain the following relation between the frequency and the amplitude of synchronous whirl: (?)v(i - A))'-- b “ (?Mk - A) e cos 0 + e2 = 82 (8-14) For a further study it will be convenient to introduce in the above equation a new variable defined by _ 1 I*03 /O 1C \ 7 Mo, 48 EI (8-15a) This leads to a simple quadratic equation, the solution of which is 7 = JUlT) \r) cos * ± cos2 ^ “ e2)]M! (8-156) By making use of Table 8-1 and the above equation, values can be found for 7 as a function of any eccentricity e. The frequency w is defined in terms of y by Eqs. (8-15). By rearranging and solving for a>, we get the following formula: CO — — ^4(^-7+w]” <»-*> 8-5. Fluid-film and Rotor Resonance. Fluid-film resonance is a con¬ sequence of the quasi-elastic properties of the oil film. The phenomenon
Hydrodynamic Instability 273 may therefore be studied by assuming that the rotor is rigid. The conditions of a rigid rotor may be simulated by setting l*/EI = 0 in Eq. (8-15a). Thus, we obtain 7 = in. By substituting the expression for y in Eq. (8-156), we obtain the following equation for w: = nLD(R/C)2f “ tM{e cos <t> ± [e2 cos2 <*> + (82 - e2))») ( ’ By considering the forces acting on the journal, it is possible to derive the frequency response characteristics for the system and predict the variation in amplitude of “fluid resonance” with speed. Referring to Fig. 8-7, we shall denote the disturbing force by the vector F2, the instantaneous position of the journal by the vector e, and the force which the fluid film exerts on the journal by F. The inertia force acting on the journal will be given by Me. If the journal is vertical and there are no other external forces acting, we can express the conditions of force equilibrium for the journal by the following vector equation: Me + F = F2 (8-18) Let md denote the amount out of balance and w the angular velocity of the rotor. The disturbing force acting on the rotor will then be given by CF = F2 = (8-19) where 0 is the phase angle of the disturbing force relative to the displace¬ ment vector e. If we assume that the journal center moves in a circular path around the bearing center, we can write e = eoe™1 where eo is constant over any complete revolution. The inertia force acting on the journal can, therefore, be expressed by Me = -Jl/c^eoe*1 (8-20) If we assume that a single bearing is taking the total load, Eq. (8-17) becomes F = MwLD (*)’ /e;<",+« (8-21)
274 Theory of Hydrodynamic Lubrication By making use of Eqs. (8-19) to (8-21) in Eq. (8-18), we obtain — Mo>2eoeiut + ^ wfiLD (Jj'j fen»t+*) = m6<a2eilat+fi) The real components in the above equation give 2 / cos <f> = mdu cos (3 and from the imaginary components we get 2 / sin 0 = mbu sin /3 — Mo>e0 -1- yLD f ^ h*D(§, By eliminating 0 between the above two equations and rearranging, we finally obtain the following equation in a>: [vLDiR/CyfY , Me0iiLD(R/Cyf cos <t> * tr[(Me,)2 - (m5)2] " "h 4tr2[(Me0)2 - (md)2 = 0 The condition for resonance is that the two roots of the above equation coincide. This will lead to the following relation: cos2 0 = i - ( \Me, and the resonance frequency will be given by _ MeofxLD(R/C)2f cos 0 - 2*l(Me,y - (mb)2] (8-22) (8-23) 8-6. Equations of Small Oscillations. We shall consider now the equations of motion for small oscillations about a position of equilibrium corresponding to some particular external, steady load. We shall employ fixed rectangular (x,y) axes as shown in Fig. 8-8. Suppose now that, under a certain external steady load, the journal is running at an eccen¬ tricity ratio co which corresponds to position Ao, which we choose as the direction of positive y. Then Eq. (8-3) yields for the steady load on the journal Pr — —ho)fr(e 0,0) Ft = Aco/*(*0,0) (8-24) Consider next a small (variable) deflec¬ tion da, dt of the journal center to A Fig. 8-8. Coordinates for small dis- and let X’Y b® the added forces on placement from equilibrium. the journal in the x,y direction result-
Hydrodynamic Instability 275 ing from this displacement. From Eq. (8-3) owing to changes in magni¬ tude of Fr, Ft we obtain (since co is constant) dFr = X dFt = X 2 da/, - — 2 daf, + w de + H] H] (8-25) df df df df where fr, ft, -A -/-*» ^7> ^7 are evaluated at € = €0, e' = 0. These forces de de de de contribute directly to X,Y, respectively. In addition, there are also contributions to X and Y due to the directional changes da given by Fr da = — Xcofr da —Ft da = — Xo>/* da where /r, ft are also evaluated at eQ, 0. Hence X = X — oj da fr — 2dof, + 00 de + ^ -u> cfa/, + 2dafr - u d» + |£ JJ (8-26) (8-27) Since /r, are positive and increasing functions of € as well as c', both frjft and the various derivatives dfr/de, dfr/de' are positive. Let the components of the displacement A0A along the x,y direction be (Fig. 8-8) and let a be measured from the y axis. Then for small £, rj, and a there results from Fig. 8-8 Eqs. (8-27) yield X = V da = C eoO _ de _ V dd = t CO Cco " e0C (dA _ 2/if 71 -L dft V €()C €0 C c de' dfr V dfr if €0 C ^ €qC de c de' C (8-28) (8-29) Now suppose that a rotor consisting of a single disk is symmetrically mounted on a shaft which is carried in two similar and symmetrically placed bearings. Let k be the shaft stiffness constant. We assume that the deflections of the shaft in each bearing, both for the steady load deflection e0 and for small further deflections, are symmetrical about the middle plane of the rotor and that steady load and the added forces X, Y on each journal are equal. The shaft segment to the left of the rotor has the forces X, Y0 + Y exerted by the lubricant on its journal and must
276 Theory of Hydrodynamic Lubrication have an equal and opposite force exerted on it by the rotor mass (which also exerts a similar force on the shaft segment to the right). Thus the shaft center deflects relative to the journal centers an amount x = — 2X giving rise to a net displacement of the rotor center z + £ y + v (8-30) (8-31) We suppose that the steady force —2F0 is an external load on the rotor (gravity). The remaining forces, 2X and 2Y, assuming that the rotor 100 dfr ifde 10 UD'-yo.T* z % Fig. 8-9. Radial stiffness for incompress¬ ible plain bearings. 0 0.1 0.2 0.3 0.4 0.5 Q6 0.7 0.8 e Fig. 8-10. Derivative of radial force with respect to velocity for incompressible plain bearings. (8-32) is balanced, can only be used up as “inertia forces.” Hence 2X = Mix + £) = -kx 2 Y = M(y + ri) = -ky and, by substituting from Eqs. (8-29) we get y _ x ( ^ r\ dft rj\ M .. kx x -x{-7JJ--^c + -dTc + d?c) = ^(-x + t) = ~2 V - \ ( vf'S _l 2fr$ ** dfr v dfr A M (a _l ky Y - + IX - ~dT C - Me) = -2iy + r,) - - T Here the functions fr)ft and their derivatives with respect to «,e' are all evaluated at the equilibrium eccentricity ratio, and for «' = 0. They are given in Figs. 8-9 to 8-12 for the incompressible cases. (8-33)
Hydrodynamic Instability 277 The differential Eqs. (8-33) are linear in the variables ^frj;xfyf and their solutions contain time as an exponential. These may be expressed in dimensionless form as err where r = coot 0)0 = /- y/M (8-34) Here wo is the critical speed of the simply supported shaft-rotor system whose mass is M and whose stiffness is k. Fig. 8-11. Tangential stiffness for incom- Fig. 8-12. Derivative of tangential force pressible plain bearings. with respect to velocity for incompress¬ ible plain bearings. From the right-hand pair of Eqs. (8-33), there results M(v wo)2 M(v wo)2 x = — k + M{vmY * 9 k + M(vw„)2'' Equations (8-33) now lead to the determinantal equation kMeoC(vo)o)2 V = - (8-35) 0)fr + 2vO)0ft + — uft + 2 VO) ofr dft , dft 2x[fc + M(^o)2] “ + at' dfr , dfr , kmC(v Wo)2 " at + ,'"0 at' + 2\[k + MOwo)2] If we now introduce the dimensionless ratio w s = — wo = 0 (8-36) (8-37) where w is the angular speed at the threshold of instability, we get Atov2 /M = wo2 Sfr + 2 Vft + — sft + 2 vfr 1 + *2 at at' Av1 | dfr j dfr at' = 0 (8-38)
278 where By writing Theory of Hydrodynamic Lubrication kC** A = r = 2/zL/?3wo Av2 1 -i- Eq. (8-38) becomes Wo sfr + 2j//* + e0r S~dt^r v — s/l + 2 k/, + f = 0 (8-39) (8-40) (8-41) wo is not, in general, equal to zero, so that the factor w02 can be divided out of Eq. (8-41). It was assumed in the derivation of Eq. (8-36) that the solutions of the equations of motion were of the form err, where v is a complex number. If the system is dynamically stable, the real part of the complex number v is negative. Conversely, if the system is dynamically unstable, the real part of v is positive. Thus, at the threshold of instability, v will be a pure imaginary number. We now solve Eq. (8-41) for the condition where v is wholly imaginary in order to obtain the value of w at the onset of instability. Considering first the imaginary part of Eq. (8-41), we have 2/, •s + * sfr + dft de' 2/r -sft dfr de' 2/, «^ + f -«/< Since v ^ 0, f (2/- + fS) + +/,S)+ 2s(f:^~fri)= 0 If s = 0, we obtain a trivial solution; for s ^ 0, we have £ = -2(/« a/,/3* ~/,a/«/a«) - (/, a/,/«V + /, a/,/a*') s 2/, + e dfr/dt' (8-42) Next considering the real part of Eq. (8-41), we have
Hydrodynamic Instability 279 Therefore ('• t.+<• f)+«(>■+• t)+f+2-’ ('• i? - *») - ° Again, for s ^ 0, we have _ -«(f/s)2 - a + ‘ a/,/*)(r/*) - (A a/,/3. + a/,/a.) } W 2(/| df,/dt - fr df,/d('\ From Eq. (8-40) we have (>2 - Av1 + f = 0 Once again, for s ^ 0, we can write „ -4(,/s)2± VM(„/sj2]2 - 4(f/«)*(*/«)* ,. itmuw (tM4) The speed a> at which instability starts to occur is now defined, since O) = SO)q. The above defined speed at which instability sets in is, in general, different from the critical speed of the shaft-rotor-bearing system. For a symmetrical, two-bearing system the critical speed may be calculated as follows: a. Shaft stiffness = k b. Lubricant film stiffness = = ^wRiR/C) df de ttC dt _ sk df “2ATt The critical speed of the system is then 2 = 1 = kJ_M_ ■"'"G'+i-s/ao' (“if),= {dpIr+ATs) (8_45) or where subscript r refers to radial stiffness. The dimensionless number A [defined in Eq. (8-39)], is a function of bearing geometry, shaft stiffness, and fluid viscosity. Calculations for
280 Theory of Hydrodynamic Lubrication the threshold of instability in which A was varied from 0.1 to 100 for 0.1 < c < 0.8 and L/D = 0.5 and 1 were performed. The values of fr, fh d/r/d«, dft/de, dfr/de', and dft/de' were obtained from the solution of the dimensionless Reynolds equation. These are given in Table 8-1 and Figs. 8-9 to 8-12. By introducing these values into Eqs. (8-39), (8-42), (8-44), and (8-45), we obtain the results of Table 8-3. The results Table 8-3. Threshold of Instability for Symmetrical Rotor Supported by Plain Journal Bearings L/D € A r s (*/«)* 1 V i s s (w<7*)r <0 too («C*)r H 0.2 0.1 - 0.9429 -0.1411 0.3756 2.6096 0.9852 2.6488 0.5 0.1 - 3.1968 -0.1252 0.3538 2.8104 0.9974 2.8177 0.8 0.1 -13.4452 -0.05543 0.2354 4.2438 0.9998 4.2446 H 0.2 100.0 - 0.9429 -0.1411 0.3756 0.06678 0.02910 2.2948 0.5 100.0 - 3.1968 -0.1252 0.3538 0.2533 0.1315 1.9262 0.8 100.0 -13.4452 -0.05543 0.2354 1.9267 0.7751 2.4857 l 0.2 0.1 - 2.6761 -0.1377 0.37 2.6763 0.9948 2.6903 0.3 0.1 - 4.324 -0.1338 0.37 2.7224 0.9971 2.7303 0.7 0.1 -18.484 -0.0846 0.29 3.4179 0.9998 3.4186 0.8 0.1 -29.2005 -0.04898 0.22 4.5169 0.9999 4.5173 l 0.2 100.0 - 2.6761 -0.1377 0.37 0.1951 0.0834 2.3420 0.3 100.0 - 4.324 -0.1338 0.37 0.3198 0.1390 2.300 0.7 100.0 -18.483 -0.0846 0.29 1.6553 0.7000 2.3647 0.8 100.0 -29.2005 -0.04898 0.22 3.1199 0.8998 3.4675 indicate that, while for low eccentricity ratios instability sets in at approx¬ imately twice the critical speed, this number increases with an increase in eccentricity ratio. Thus, the onset of instability for eccentricity ratios of 0.8 is about four times the critical speed. This conclusion agrees with observations which show that stability increases with an increase in eccentricity ratio and also that instability may occur even at high eccentricity ratios. The number (1 /i)(v/s) shown in Table 8-3 (where i = y/ — l) repre¬ sents the ratio of the frequency of the oscillation of the shaft center to the running frequency of the shaft, calculated at the onset of instabil¬ ity. Note that this ratio is always below 0.5 and is independent of the magnitude of A. The analysis described above is equally applicable to compressible and incompressible fluids. It requires only that the force and gradients be evaluated in each case from the applicable Reynolds equation. For the compressible fluids the forces and their derivatives with respect to dis-
Hydrodynamic Instability 281 placement and velocity are given in Chap. 5 where fr = cos <f>/S and ft = sin 4/S. 8-7. Equations of Motion for Large Displacements. We now consider the general equations of motion of the journal center and the rotor center, without making the assumption used in the preceding section that this motion is close to the steady-state position. Suppose that under a steady load on the rotor, such as gravity, the journal center is deflected to Qo, corresponding to c = €o, and that the coordinate axes are chosen as in Fig. 8-13 so that the y axis passes through Qo. Then the coordinates of Qo are i = o = Ce o Fig. 8-13. Coordinates for /o ac\ large displacements from equi- (8-46) Hbrium. The forces exerted by the lubricant on the journal at each bearing (half the applied steady load) are given by Fx o = Xw/<(€o,0) Fvo = — Xco/r(€ o,0) (8-47) These forces also cause a steady deflection QoAq of the shaft and rotor centers equal to (8-48) Xo = —2FX( ~k~ = k Consider now a general motion of the journal center to the point Q specified by the polar coordinates (eC,a). The journal center is now at £ = C* sin a rj = Ce cos a * (8-49) The forces on the journal in the radial and transversal directions are given in Eqs. (8-4); along the x, y directions they are Fx = X(w — 2a)[— /r(€,c') sin a + /<(€,e') cos a] Fv = — X(w — 2d)[/r(€,c') cos a +/*(c,e') sin a] The shaft-center deflection is -2 Fx — 2X k —2F„i (8-50) y = k — 2F, (co — 2a) (—fr sin a + /* cos a) 2X (8-51) - = (co — 2a) (fr cos a + ft sin a) and the net deflections of the rotor center are
282 Theory of Hydrodynamic Lubrication By subtracting the steady force components equation (8-47) from Eqs. (8-50) we obtain Fz - Fz0 = -Aa>/<(e0,0) + A(a> - 2d)(-/r sin a + ft cos a) Fv — Fyo = Aw/r(«o,0) — A(a> — 2a) (fr cos a + ft sin a) These forces can only be used up in producing acceleration of the rotor mass. Thus, y (* + £') = F. - F,„ ^(y + f,) =F„- F„o (8-54) or more explicitly, + [(o’ - 2a)(—/, sin a + f, cos or)] ,, n,,MCd* . . = «/i(to,0) + 2x^2 (« Sln “) (8-55) ,, MC d2 , , = w/r(«0,0) - JJjj (« C0S «) If we consider the case of a massless rotor, the terms involving M can be deleted from Eqs. (8-55), which then reduce to ^1 - ^ l—M*,*') sin a + /i(e,e') cos a] = ft(eo,0) ^1 - ^ [/r(M') cos a + /«(€,€') sin a] = fr(e0,0) (8-56) These are differential equations from which a, e can be solved as func¬ tions of €, a. The steady-state solution is given by e = e0, a = 0. Equa¬ tions (8-56) may be transformed by multiplying first by (— sin a, cos a) and adding, and then by (cos a, sin a) and adding. There results (l - —)/r(M') = -/t(*o,0) sin a + /r(co,0) cos a / 2a\ (8_57) fl - = /*(«o,0) COS Q! + /r(eo,0) sin a From Fig. 8-14 we have the following /(«,*') = V7?Tf? yp = tan-1 ^ Jr fr=f cos \p ft = / sin \p fo and are the values of / and yp at the positions of equilibrium, or fo = /(e0|0) = ^(co,0)
//ydrodynamic Instability 283 Fig. 8-14. Representation of resultant force. By taking the square root of the sum of squares of Eqs. (8-57) and their ratio, one obtains the following equivalent form for these equations: (* " =f° (8-58) tan \f/ = tan (a + ^0) Hence \f/ = a + (8-59) Let t = vt; then Eqs. (8-58) can be replaced by (1 - 2= /(«o,0) = fo (8-00) By means of plots of / and \p in the e — I plane, such as those of Figs. 8-15 and 8-16, the locus of a shaft center initially displaced from its equilibrium position (e0,^o) to some other position («,^) can be obtained numerically. The equations required for finding this locus and a description of the procedure follow: From Eq. (8-59) we have a = \f/ — \f/ o (8-01) and from Eq. (8-60) we have = 3 - \j (8-C2) For a small dimensionless time increment Ar = r„+i — r„ we can write €T-rn + | ~ + *T—Tn At (8-63) and «r=rB+l « otr=Tn + oj-r, Ar (8-64) Now, consider a shaft center whose static equilibrium position is (€0,0) and let it be displaced to a position (€T=To, aT-T0) at the instant r = r0. Its locus can then be traced as follows (see Fig. 8-17):
284 Theory of Hydrodynamic Lubrication -2.4 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.1 0.2 0.3 0.4 Q5 0.6 0.7 Fig. 8-15. Contour lines of / in « — i' plane (L/D = 1). Fig. 8-16. Contour lines of f in e — i plane {L/D = 1). 1. From Figs. 8-15 and 8-16 obtain the dimensionless force fo and the angle which correspond to the position of static equilibrium (e = co, i' = 0). 2. From Eq. (8-61) calculate \l/T-r0 = + Or-r0- 3. Locate the point (€rDT0,^r_r0) in Fig. 8-16 and hence read off i'_To. 4. Locate the point (eT-T„eT^0) in Fig. 8-15 and read off /T^0. 5. By means of Eq. (8-62) calculate 2 2 /o 6. For a small time increment Ar = n — ro, calculate from Eq. (8-64) T, Otr—To ~F T0 Ar 7. From Eq. (8-61) calculate $0 "F Otrmi^i 8. From Eq. (8-5) obtain = [e'd - 2<*')U, 9. From Eq. (8-63) calculate ^T-rri —To ~F ^f-»T0 At The new location of the shaft center (€T_T„aT^l) is now defined. The procedure is repeated for a new time interval Ar = n — ri, and the
Hydrodynamic Instability 285 location of the shaft center at time r = 7-2 is similarly obtained. Step¬ wise repetition of this procedure will then define the complete locus of the shaft center. Figure 8-176, c, and d shows three possible loci. In Fig. 8-176 the shaft center moves back to the position of static equilibrium so that the condition is stable. In Fig. 8-17c the shaft center describes .(£ bearing 9 O a id) ic) id) Fig. 8-17. Locus of shaft center, massless rotor solution: (a) procedure for determining locus of shaft; (6) stable condition, e.g., acceleration; (c) stable oscillation, e.g., syn¬ chronous whirl; (d) unstable oscillation, e.g., half-frequency whirl, resonant whip. a closed loop. This is a condition of stable oscillation. The period of these oscillations is given by (l/co)(rfinai — r0). In Fig. 8-17d the locus of the shaft center spirals out and the oscillations are unstable. SOURCES 1. Newkirk, B. L.: Varieties of Shaft Disturbances Due to Fluid Films in Journal Bearings, Trans. ASME, vol. 78, pp. 985-988, 1956. 2. Sternlicht, B.: Elastic and Damping Properties of Cylindrical Journal Bearings, Trans. ASME, Ser. D, vol. 81, June, 1959. 3. Poritsky, H.: Contribution of the Theory of Oil Whip, Trans. ASME, vol. 75, pp. 1153-1161, 1953. 4. Boeker, G. F., and B. Sternlicht: Investigation of Translatory Fluid Whirl in Vertical Machines, Trans. ASME, vol. 78, pp. 13-20, 1956. 5. Robertson, D.: Whirling of a Journal in a Sleeve Bearing, Phil. Mag. ser. 7, vol. 15, pp. 113-130, January, 1933. 6. Hori, Jukin: A Theory of Oil Whip, J. Appl. Mechanics, vol. 26, pp. 189-198, June, 1959. 7. Sternlicht, B., H. Poritsky, and E. Arwas: Dynamic Stability Aspects of Com¬ pressible and Incompressible Cylindrical Journal Bearings, First Intern. Gas Bearing Symposium, 1959. (Sponsored by ONR.)
CHAPTER 9 ADIABATIC SOLUTIONS INTRODUCTION In the preceding chapters dealing with hydrodynamic bearings it was assumed that isothermal conditions prevail. Yet while the fluid is being sheared, work is done on it, which raises the lubricant temperature. In the case of incompressible fluids this means a lowering of the fluid viscosity; the reverse is true in compressible fluid lubrication. The conservation of energy principle requires that the internal energy stored in the lubricant be equal to the work done on the lubricant by the viscous forces less the heat conducted away from it. Isothermal flow implies that all the work done on the lubricant is conducted away as heat. However, in most cases almost all of the work done on the lubricant is stored in it as internal energy; that is, the flow is essentially adiabatic. The actual process will lie somewhere between the isothermal and adia¬ batic conditions, and it is thus appropriate to establish both limits. This chapter deals with the latter thermodynamic path. Energy considerations in lubrication are important, for they affect the load-carrying capacity, the film thickness, and all other performance characteristics of the bearing. Satisfactory operation of bearings depends to a very large extent on maximum bearing temperature from the standpoint of material choice; it also depends on thermal gradients which in turn influence the bearing distortion and thus film profile. It becomes evident that only through the simultaneous solution of the Reynolds and energy equations can we obtain such parameters as maxi¬ mum temperature and an accurate minimum film thickness, the two parameters which are generally employed as criteria for bearing failure. One of the difficulties encountered in dealing with energy relations stems from the fact that there are no exact analytical expressions for viscosity as a function of temperature and pressure. Numerous empirical expressions of m = f(p,T) have been developed, and some of them are given below. If the lubricant viscosity at a pressure pi and a temperature Ti is represented by /ui, the viscosity at an arbitrary pressure p and tempera¬ ture T can be represented by 280
Adiabatic Solutions 287 (P - Pi) + A T Ar a T i + a 1 (9-la) This equation has a theoretical basis in the kinetic theory of liquids besides being verifiable experimentally over a useful range. The pressure and temperature scales can be so chosen that pi and Ti are both zero. Equation (9-la) then reduces to The temperature rise likely to be encountered in practical problems is small enough that for most lubricants the second term in the above equa¬ tion is nearly linear in T. Therefore an approximation to the above equation is the widely used expression Where the pressure and temperature range is large, the following expression is often used: Where viscosity is only a function of temperature, Eq. (9-Id) or (9-lc) is used: where a, 0, and y are constants whose values differ in the above five expressions. The density of most lubricating oils varies linearly with temperature, decreasing about 2 per cent for a 50°F temperature rise. It may be readily checked that for a reasonable temperature range the following formulas are very close approximations to this relationship. where pi is the density of the lubricant at 7\. The lubricant enthalpy is given by p = pi^e (9-16) (9-lc) 7 (9-Id) i + arf+ ar (9-lc) p = Pie~XT p = Pl[l + a(T - T\)\ (9-2 a) (9-26) (9-3a) In the developments that follow in this chapter the internal energy is assumed to be independent of pressure and linearly related to tempera¬
288 Theory of Hydrodynamic Lubrication ture. The internal energy equation is then 6 = 6\ cT (9-36) where e\ is the internal energy of the lubricant at some Ti and c is the specific heat. Since the energy equation (1-26) developed in Chap. 1 is used exten¬ sively throughout this chapter, it is for convenience rewritten below in slightly different form: For compressible fluids an additional term + \2h(dp/dt) appears on the right-hand side of Eq. (9-6). For infinitely long bearings the above equation reduces to The mechanism of film formation due to changes in film temperature will be presented by investigating the one-dimensional thrust bearings because the question of film divergence does not arise, and the boundary conditions can be stated in simple form. By employing the dimensionless parameters For infinitely long bearings the above equation reduces to Likewise the Reynolds equation (1-66) is given as = 6(1/, - Vt) L (ph) + 12pV (9-6) (9-7) These equations are employed throughout this chapter. THE THERMAL WEDGE Ml - _ pB rp _ TpgcB P 1 ...rr 9 P = - Pi 6mi U mi U Equations (9-5) and (9-7) can be reduced to IK '-j©]-° <«•>
Adiabatic Solutions 289 These can be integrated to give dx frh* H 2 ^ 4 V M / J = -?■ (l - (9-96) dx h*\ ph) where Poho is the constant of integration. Now for successful lubrication the pressure must rise to a maximum and fall again, that is, dp/dx must vanish at some point in the film. The equation shows that evanescence of dp/dx cannot be achieved by varia¬ tion of p alone, since viscosity is a monotonic function, but that it can be achieved by suitable decrease of p> of h, or of their product. Quali¬ tatively, the effect is the same for variation of p or of h, but quantitatively it will be different, since l/h2 occurs outside the bracket. A decrease of h will produce a greater total change in dp/dx, so varying h is likely to be more effective than varying p. If p and h both vary in such a way that ph is always >pohoj then again film lubrication is impossible. Now, the pressure in a bearing increases manyfold but the temperature increases only a fewfold, so that: 1. For a gas p( p/T) will increase. So it seems unlikely that it will be possible to run parallel face bearings with gas as a lubricant, and it will be noticed that in general the variation of h will have to be considerable to counterbalance the increase in p. 2. For a liquid the increase of p with p is usually small compared with its decrease with T, and the over-all variation of p may prove sufficient in some cases for successful film lubrication of parallel face bearings. Since for all real fluids p and p increase or decrease together, an increase of pressure inhibits the formation of a film and the only useful variation of p is the decrease with temperature. The boundary conditions are x = 0, p = 0, T = Th p = p = 1; x = 1, p = 0 and the variation of h, p, and p is given by E-'-O-TT1) ' - 1 - MT - n "(0 where h\y /i2, A, and are known a priori. With an assumed pohoj Eq. (9-9a) can be integrated immediately to give the dependence of T on x. From this relationship the dependence of p and p on f can be replaced by a dependence on x, so that Eq. (9-9b) becomes integrable. If the assumed poho is correct, then p will vanish for x = 1; if it does not, then another value of poho is taken and the process is repeated as often as is necessary to make p vanish for x = 1 to the required order of accuracy.
290 Theory of Hydrodynamic Lubrication Table 9-1. Hydrodynamic Equations of the Geometric and Thermal Wedges Equations Geometric wedge — = - (\ - dx ~ h>\ h) dT=±\l _*h_o dx h0h [ 2 h 4\h) Thermal wedge dp = J_ (. _ po\ dx h.* \ p / = AT. _ :lb + ?fiLY dx h2* [ 2 p 4 \ p / Boundary conditions x = 0, h = h\, p = 0, T = T\ x = ], h = h2, p = 0, T = T2 x = 0, p = 1, p = 0, T = f, x = 1, p = p2, p = 0, T = f, Definition of a hjL. MH Law of vari¬ ation h = ah p = 1 - (a - l)x h0 or po :2ah2/(a + 1) (a — l)/a ln c (£, — h)(h — h2) (a2 - \)(h/h*j~ [-T (1 - P) In (1 /P) ln a p (a 6 f 2(a - 1)1 \a — 1 In a) AT o(a + 1)i 2 In a — 3 6 <^’-4 a(ln a)2 Load criterion Drag criterion F = load per unit area vxU/B frictional drag per unit area (h ~B ftp* nxU/B dx rp, . A temperature rise through bearing , - _ Temperature criterion AT = - k22 = (T2 — Tx)ho* uxU/gpxcB Solutions obtained from Eqs. (9-9) cannot, of course, pretend to repre¬ sent all that goes on in an actual bearing. Besides any approximation involved in the assumptions which lead to Eqs. (9-8), the simplest pos¬ sible equation of state is used, and the law of viscosity variation selected is a compromise between the actual viscosity variation and an expression
Adiabatic Solutions 291 Table 9-2. Effect of the Geometric and Thermal Wedges on the Performance of Plane Sliders a ht/B a 0 ... ..p .... F F/P AT 10"» 0 0 47 X 10"3 0.95 20 1.81 '% lO"8 0 -1.5 32 X lO"3 0.63 20 1.18 10-» 0 -3 24 X 10“3 0.48 20 0.91 1.25 10"3 0 0 88 X 10"3 0.90 10 1.63 1.27 10"3 10“3 0 95 X 10"3 0.90 9.5 1.65 1.25 10"3 0 -1.5 61 X 10"3 0.61 10 1.09 1.26 10“3 10"3 -1.5 64 X 10"3 0.61 9.6 1.10 1.25 10“3 0 -3 48 X 10-3 0.48 10 0.84 1.26 10"3 10-3 -3 49 X 10"3 0.48 9.6 0.85 1.5 10~3 0 0 131 X 10"3 0.84 6.4 1.41 1.5 10"3 0 -1.5 95 X 10"3 0.60 6.3 0.98 1.5 10“3 0 -3 75 X lO"3 0.47 6.3 0.77 1.75 10"3 0 0 151 X 10"3 0.80 5.3 1.26 1.76 10~3 0 0 154 X 10~3 0.80 5.2 1.27 1.75 10"3 0 -1.5 112 X 10~3 0.58 5.2 0.89 1.76 10"3 10"3 -1.5 115 X 10"3 0.58 5.1 0.90 1.75 10"3 0 -3 92 X lO"3 0.47 5.1 0.71 2 10~3 0 0 159 X 10"3 0.77 4.9 1.16 2 10~3 0 -1.5 122 X 10“3 0.57 4.7 0.84 2.01 10“3 10~3 -1.5 123 X 10"’ 0.57 4.7 0.84 2 10”3 0 -3 101 X 10"3 0.47 4.6 0.67 1.02 10"3 10"3 0 11 X 10~3 1.00 92 2.02 1.01 10"3 10“3 -1.5 4 X 10~3 0.64 147 1.29 1.01 10“3 10"3 -3 3 X 10"3 0.48 196 0.97 1.06 6.32 X 10~* 10'3 0 28 X 10“3 1.00 36 2.06 1.02 -v/40 X 10"4 10"3 -1.5 6 X 10"3 0.47 79 0.95 1.02 V^O X 10"4 10"3 -3 3 X 10-3 0.32 120 0.64 1.13 4.47 X 10"4 10"3 0 60 X 10“3 1.00 16.7 2.13 1.04 y/20 X 10"4 10~3 -1.5 7 X 10-3 0.35 52 0.72 1.02 y/20 X 10"4 10"3 -3 3 X 10-3 0.22 88 0.44 1.33 3.16 X 10"4 10'3 0 142 X 10“3 1.00 7.0 2.35 1.06 V10 X 10"4 10"3 -1.5 7 X 10-3 0.26 36 0.53 1.03 Vio X 10“4 10"3 -3 2 X 10"3 0.15 67 0.30 a is the coefficient of cubical expansion of the oil with temperature: p = p,U + a(T - TO] 0 is the index of T in a power-law variation of viscosity with temperature: a is the ratio of inlet to minimum wedge thickness _ h ° ~ h, + 1 - A' AT where K is written for ap\UB/gpxch22
292 Theory of Hydrodynamic Lubrication that is analytically simple. Nevertheless, useful information can be obtained or inferred from a set of solutions with h%9 h2j A, and 0 selected to cover the range of variation actually encountered in practice. Let us consider solutions of Eq. (9-96) for the two simple cases, namely, h a linear function of x, the pure geometric wedge, and p a linear function of Tf the pure thermal wedge. The results are given in Table 9-1, where because of its complexity the solution of the energy equation is not given. This solution does, however, show that, for constant viscosity and light loads at any rate, T is approximately a linear function of x. The analysis can thus be considerably simplified by replacing the Unear dependence of p on T by a linear dependence of p on 5; accordingly, the formulas for the thermal wedge in the table have been derived for p « x. The definition of a also requires clarification when both geometric and thermal wedges are present. The preceding discussion shows that l/p2 plays the same part for the thermal wedge as h\/h2 does for the geometric. The simplest procedure is to define a as the ratio From this definition the value of a has been calculated. The solutions of Eqs. (9-9) under the conditions stated above are given in Table 9-2. 9-1. Parallel SUder with p = f(T) and p = f(p,T). The Reynolds equation (9-7) and the energy equation (9-5) with the density, viscosity, and internal energy given by Eqs. (9-2a), (9-16), and (9-36), respectively, represent a system of simultaneous equations in pressure and temperature as the two unknowns. Equation (9-7) after integration becomes Inlet film thickness “Effective film thickness” at point of closest approach Thus, when both geometrical and thermal wedges are present, h2 P2 Ki A(T2-Ti) h2 1 - A(T2 - fi) hi (A/h22)AT h2 1 - (A/h22) AT ONE-DIMENSIONAL SOLUTIONS (9-10)
Adiabatic Solutions 293 Upon substitution of Eq. (9-10) and the expression for internal energy into the energy equation (9-5) and rearrangement, there results idra 2 U p dx h2cpog From the density equation (9-2a) we can write — = eX(r_T,) = 1 + X(T - T„) + % (T - To)2 + • • • P * In all cases of practical importance, only the first two terms of this series need be considered, so that Eqs. (9-10) and (9-11) reduce to 1 dp _ 6U./rp , IdT _ 2U — -j— — , 2 ^ (T ^ o) and -■ ~j — p dx hr p dx h2cpog respectively. Upon introduction of viscosity equation (9-16) into the above two equations and rearrangement, there results, respectively, d(-e~“P) = - To) (9-12) dx h2cpog The pressures usually encountered in parallel-surface slider bearing lubrication are such that their variation is small. If eap is replaced by ki, a constant, Eq. (9-13) can be readily integrated, yielding the following temperature equation: T = ^ln (1 + kjex) with Jfc = (9-14) p n£cpog where it has been noted that T = 0 when x = 0. Upon substitution of this temperature equation into Eq. (9-12) and integration, there results the following pressure equation where it has been noted that p = po when x = xq. If ki is set equal to unity, the minimum value of eap, the temperature distribution given by Eq. (9-14), will bound the actual temperature distribution from below. Since pressure varies directly with viscosity and viscosity varies inversely with temperature, the pressure distribution given by Eq. (9-15) for this value of ki will bound the actual pressure distribution from above. If k\ is now set equal to eapo, the temperature distribution given by Eq. (9-14) will bound the actual temperature
294 Theory of Hydrodynamic Lubrication distribution from above. The pressure distribution given by Eq. (9-15) for this value of k\ will bound the actual pressure distribution from below. Thus, in this way, close upper and lower bounds can be established for the pressure and temperature in the oil film. The three quantities x0, po, and T0 are still unknown, but since Eq. (9-14) relates T and x, there are really only two unknowns. There are two boundary conditions available for the determination of these unknowns, namely, p = 0 when x = 0 and x = B. By eliminating x between Eqs. (9-14) and (9-15), we have e-ao-p.) = I + 3aXcpoff e„p^T _ To)2 (9_16) 2k i For both this equation and the boundary conditions to be satisfied, it i£ easily seen that it is necessary for AT, the temperature rise from inlet to outlet, to be twice T0. With this relationship Eq. (9-14), when solved for AT, yields the following equation for T0: T° = ^ln (1 + k\k&) The above equation is implicit in T0, since T0 is contained in k. Once T0 is known, Eq. (9-14) written in the form provides the value of x0. Equation (9-16) applied to either end of the slider yields ZaXcpig 2ki With T0, xo, and p0 known, the temperature and pressure equations (9-14) and (9-15) are now complete. 9-2. Step Slider with p = /(x), /u = /(p,T). Starting with the Reynolds equation (9-7) and integrating it leads to (9-17) where h0 is the constant of integration. Upon substitution of the above equation into the energy equation (9-5) there results f-7JMr.[4»’-67*J‘ + 3(>)1 (W8) We shall employ Eq. (9-16) for viscosity variation and assume that the density varies exponentially with distance along the slider; that is p = pjg-xr/B (9_19)
Adiabatic Solutions 295 This variation is almost linear, since X is small. The reasonableness of this assumption is verified later. The constant X must be related to the outlet temperature T2 by X S ln ^ = 7^2 (9-20) P2 with y the coefficient of cubical expansion of the lubricant. By differentiating Eq. (9-17) with respect to x and combining it with Eqs. (9-16) to (9-19), we obtain d /mi\ r „ hoe*z,B - h , 4/i2 - m0e^B + 3/i0Vx"* _ [ _) = 6a Ufx i ^ + dx\nj /j,3 1 gpiC h0h3 If we introduce the following dimensionless notation (9-21) Pi r M — h n n i and consider the stepped slider of Fig. 9-1, the above equation becomes (9-22) d/Z „ heu — s „ 4s2 — 6s/ieXI + 3h*e2Xx -jr = /Vl o h A2 "7 dx s3 ^ where 6a/xi UB K2 = Eq. (9-22) integrates to n = fir, + K: X(eXi — ex,J) — Xs(x — 77) Xs3 where , 4Xs2(x - v) - - ex”) + V2h2(e^ - e2X”) ,n oox + K2 (9'23) 77 = 0 0 < x < 6 7} - b b < x < 1
296 Theory of Hydrodynamic Lubrication Note that ft = 1. If the fact that the pressure and temperature are continuous at the step is taken into account, we have that ft = ft where subscript h corresponds to position x = b. Upon replacing the exponentials in Eq. (9-23) by their power-series expansions and retaining only terms in the first power of X (since X is quite small), there results £ = m -|- Xd,(x2 — ri2) with m — ft + c,(z — rj) , „ h — s v 4s2 — 6s/i + 3h2 where c, = K\ —5 h K2 = S3 s*h J TS 1 f 3) d'~ Kl2? + K* ? Note that m is the value of fl when density variation is neglected and that ft = mi + \djb2 where subscript 0 corresponds to boundary condition *7 = 0. Equation (9-17) for the stepped slider can now be rewritten as dp _ 6mi UB s h (1 -f- \x) dx hi2 53[m + Xd,(z2 — 172)] Equation (9-24) must be integrated with respect to x to determine the pressure distribution. The constant h is determined by the condition that the pressure vanish at the inlet and outlet ends of the slider and be continuous at the step. With this value of h, the pressure distribution may be integrated with respect to x to determine the load capacity. The temperature rise through the bearing may then be calculated and the constant X adjusted to satisfy Eq. (9-20). This direct procedure is tedious and quite time-consuming. A relatively simple iteration pro¬ cedure for finding the load capacity is described below; the first step of the procedure proves to be sufficient for almost all cases. If the density variation is neglected (that is, X is taken as zero), it is apparent that, since there is no f4 density-wedge/* the resulting load capacity will bound the actual load capacity from below. For this case Eq. (9-23) becomes dp _ 6niUB s — h dx hi2 s3m By integrating the above equation with respect to x, we obtain
Adiabatic Solutions 297 The condition that the pressure be continuous at the step requires that a — h . 1 — h. mi —:— ln mi = ln — a3co Cb m2 The constant h can be determined from the above equation. Integrat¬ ing Eqs. (9-25) with respect to x yields for the load capacity w _ GmULB* - [mi ln ^ + (mi - m2)l Cb i mi J (9-26) Thus, if the geometrical configuration of the slider (£,a,6), the operat¬ ing conditions (hi,U), and the lubricant properties (/z,a,7,pi,c) are known, Eq. (9-26) gives a lower bound to the load once h is determined. From the definition of ra2 it follows that the temperature experienced by the lubricant in flowing adiabatically through the stepped slider bearing is to a first approximation given by T2 = - ln m2 y If the temperature as given by the above equation is substituted into Eq. (9-20) to determine a X, a second approximation to the load capacity can be obtained by substituting this value of X into Eq. (9-24), integrating this equation to obtain the pressure distribution, determining h by requiring continuity of pressure at the step, and with this h integrat¬ ing the pressure distribution for the load capacity. If greater accuracy is desired, the cycle of operations can be repeated. 9-3. Exponential Slider with n = f(p,T). The Reynolds equation (9-7) integrated and rearranged reduces to s-wo-*) By substitution of Eq. (9-27) into the energy equation (9-5), there results dT 2 nU dx gpcho2 If we let h0/h = h and assume the film thickness to vary with distance along the slider according to the equation h = hie~bx, Eqs. (9-27) and (9-28) become 5?-§?«■-*> <9-29> and ~ (3£2 - Qh + 4) (9-30) dh bgpcho respectively.
298 Theory of IIydrodynamic Lubrication Upon introducing the viscosity equation (9-16) into Eqs. (9-29) and (9-30), we obtain = &ua^fi_ 1} Mi dh bh<? and - e-“» = 2t/^ (3/i2 - 6h + 4) mi dh bgpcho The sum of the above two equations is recognized as (d/dh)p~l. There¬ fore the reciprocal of viscosity at any point in the film is given by r* d(iu-*) dh This leads to f ■dh + u? [(" + £) - *'•> -1 (“ +1) <S’ - S''> + «(*-«,)] (9-31) By substituting Eq. (9-31) into Eqs. (9-29) and (9-30), we obtain the following pressure and temperature distribution equations: 6/xiU h( 1 - h) dh T- k1 bhj h, 1 + (2niU/bhj) Y (9-32) (9-33) bgpcho2 Jhl 1 + (2piU/bh02)Y where Y is the expression within the brackets in Eq. (9-31). One boundary condition has not yet been used, namely, p = O.when h = hi. From this condition h0 can be determined, thus making the pressure and temperature equations complete. Since the integrand in Eq. (9-32) contains a cubic h in the denominator, it is impractical to determine ho explicitly, and a trial-and-error process will prove more convenient. Once h0 is known, the integrations indicated by Eqs. (9-32) and (9-33) can be carried out to give the pressure and temperature at any point in the film. FINITE SOLUTIONS The energy and momentum equation developed in Chap. 1 is rewritten in polar coordinates to give +;s[(^+£)*(- £8+?)]--(£8+*?) <»*>
Adiabatic Solutions To put the above equation in dimensionless form, let 299 r = Ri h2 6 A = — mi P = 12tt.VVi T = KT where where K = gPCyJ Substituting the above dimensionless parameter in Eq. (9-34) gives us T + V N frfc ?P\ , j>_ A* dp\ _ _2 Ai f2 [ df\ji df) d5\pde) d9\ + a(-f + P) + (-*1 dl + A d(-T + P> \ p df) df \ jifi de ) de = h (9-35) desk The first part of Eq. (9-35) is identically equal to zero and is the Reynolds equation. The remainder of Eq. (9-35) is the energy equation (1-26) in polar coordinates. Next, the dimensionless Reynolds and energy equa¬ tions can be transformed into a set of difference equations (refer to Chap. 4) as follows: (? Pt+i. A Af2 y + (tI Af2 / ^ A3| li. Pt.y+A y+>4 Ad2 / ( rhz ir^3 \ 1 . /£3| i ) — It H—_- t-Bl> M 1 VM It.y+w + ~_ rp It. Ad2 i. (9-36) rhe A2 / ’ V Ad / /rfr I rh* \ J /£* I fi* I \ |t + M*. j M i — Vz.j/ Af2 vAl«,j + W fp\i,j-Yi) Ad2 energy equation in dimensionless form reduces to Air2 . h* / dp V i /dpVI — /fr fr3 dp\ dT /t3 dp df 3^ |A^/ VV J V pf2 dd/ dd M df df herring to Fig. 9-2, the above equation reduces to a difference equation: 2 _j_*L3| - ft.i-A2 | + ft-i.A2l ■« *LlV 2fA* ) 2Af )\_ = [a I — — I (f>i i+' ~ /r..j+1 — tA L L- ^l,A 2A9 /J V 2 A § / A I /ft+l. j ~ P<-1. A /r,+ l. j ~ 7*1-1. A ~H,A '2Af A 2 /
300 Theory of Hydrodynamic Lubrication Fig. 9-2. Indices used in pressure and temperature calculations. Solving next for fit gives ;2' Ti.j+i = 2A0 [T'! + *!| I CP<-^ - P Wu g I.J 2 f Ae a* l //>.+1 + M l.j _ I (T. A] ij pr2 lt.j V 2 Ad / J r+i.y ~t~ Pt-i. A /Tj+i'j ft-i,A 2 Af / \ 2 Af / L-l _ il I /Pi.i+1 - P.-.y-iM li.y Mr4A 2 AS )\ [■ + '_£.| (*».- ij Mr- |itj \ t ~ Pt. y-A 1 y-i 2 Ad / J 2 Ad k; pr2 I*,; 1 ~ 2 Ad )\ (9-37)
Adiabatic Solutions 301 The steps for a simultaneous solution of the Reynolds and energy equations may be performed in the following manner: 1. The value of the film thickness at every point is determined. 2. pi,j is assumed equal to zero and the known value of inlet tempera¬ ture is assigned to T\. 3. The values of 7\,y are then determined at every point from Eq. (9-37), which provides us with ft,,. 4. Having the values of ft,,-, /i*-.,-, and ft.,-, the first approximation of the pressure field is determined from Eq. (9-36) and improved by iteration. 0.4 0.6 0.8 ft rodions Fio. 9-3. Adiabatic load capacity of ta- pered-land bearing. L/R% = 0.5; — L/Ri = 0.56; - 2.93 X 10"® lb- sec/in.2; n at 100 = 4.15 X 10“* lb-eec/ in.2; n at 210 = 0.63 X 10"® lb-sec/in.2 0.6 0.8 ft rodions Fig. 9-4. Adiabatic pressure gradient out¬ flow for tapered-land bearing. L/Rt = 0.5; — L/Rt - 0.56. 5. The value of the pressure field thus obtained is used to recalcu¬ late the temperature distribution from which a new set of ft,,- values is determined. 6. A second approximation of the pressure field is now obtained. This cycle of pressure and temperature iterations is continued until the error, which is the difference between successive values of the pressure field, falls within the limit prescribed in Eq. (9-37). 7. The final value of the pressure field is then used to compute the final value of the temperature field. With the pressure and temperature fields known, the bearing perform¬ ance calculations can be carried out. The above procedure, steps 1 to 7,
302 Theory of Hydrodynamic LvJbricalion yields an adiabatic solution, i.e., one which neglects heat conductance through the bearing and shaft. An isothermal solution may be obtained by an iterative solution of Eq. (9-36) only, by a method described in Chap. 4. Following are several solutions obtained by considering adia¬ batic conditions for sector thrust bearings with two film shapes. Fig. 9-5. Adiabatic load capacity of Michell film shape. Dashed line, L/Rt = 0.5; —L/R« = 0.56; C,/A* x = 1, O = 1.5, • = 2. j9, rodions Fig. 9-6. Adiabatic pressure gradient out¬ flow for Michell film shape. C\/h» x = 1, O = 1.5, • = 2. Tapered-land Film Shape. The following equation represents a typical film shape of a tapered-land bearing. A = Ci(l^»)J3 - t(Ar)/L] + hi hi hi where Ci is a constant, i is running index in radial direction, and Ar is mesh size in radial direction. Figures 9-3 and 9-4 illustrate, in dimen¬ sionless form, the force and “pressure gradient outflow” characteristics for a tapered-land bearing as a function of pad angle 0.
Adiabatic Solutions 303 Michell Film, Shape. The film shape often referred to as a Michell bearing may be expressed in dimensionless form by A = 1 + £> (1 - 9) fl 2 H>2 (9-39) where C\ = 5. Figures 9-5 and 9-6 illustrate in dimensionless form the force and pres¬ sure gradient outflow characteristics of a Michell bearing as a function of pad angle. Table 9-3 shows the dimensionless “pressure gradient Table 9-3. Power Loss in Adiabatic Thrust Bearings h-t, in. Tapered-land: 0.0010 0.0015 0.0020 0.0030 0.0050 Michell: 0.0010 0.0020 0.0050 0.0010 0.0020 0.0050 0.0010 0.0050 in. 0.0030 0.0030 0.0030 0.0030 0.0030 0.0010 0.0020 0.0050 0.0015 0.0030 0.0075 0.0020 0.0100 (hpip + Aps + hi) X 10* L/Ri = 0.5168 L/Rt = 0.5000 0 = 0.69 0 = 0.76 1.88 2.18 2.40 2.55 2.52 0.93 1.55 2.20 1.77 2.79 3.81 2.65 5.45 2.15 2.52 2.80 3.05 3.10 1.88 2.72 2.00 3.38 4.53 3.13 6.50 0 = 0.84 2.30 2.75 3.02 3.32 3.38 2.00 2.97 2.18 3.60 4.81 3.29 7.17 2.47 2.96 3.26 3.59 3.66 2.14 3.21 2.33 3.88 5.48 3.55 7.77 horsepower” loss as a function of bearing geometry for Pi = 2.93 X 10-8 lb-sec/in.2 The total dimensionless flow out of a bearing is where Q = Q 2» + Q2p Qz Q4 231Q and 7 60a- NRiLh, while the total dimensionless horsepower loss is 2 Rt where hp = hpiV -J- hp2P + hp$ -f- hp4 ^ (horsepower) = q AT (9-40) (9-41)
304 Theory of Hydrodynamic Lubrication Figure 9-7 is included, for it compares the load-carrying capacity, using several methods of analysis, for the tapered-land and Michell bearings, respectively. With isothermal assumptions, the two bearing profiles carry approximately the same load. With adiabatic assump¬ tions, the tapered load bearing is somewhat better. Even though the isothermal solutions employed average lubricant viscosity in their calcu- 1 1 1 -O.OO6V-7 0.000” -0.( )06"'— Jom 3" 1 Pi y 8=1/^ k UfrZ/fT' 1 y 8=5/6 1 0 0.001 0.002 0.003 0.004 0.005 0.006 hlt in. [a) Tapered land film shape • Isothermal (exponential solution ref. 5) pa¥g - 2.04 x 10'6 lb sec/in.2 o Isothermal (iterative solution ref.7) /i^ = 2.04 x 10“6 lb sec/in.2 + Adiabatic (iterative solution ref.7) /*, = 2.93 x 10“6lb sec/in.2 x Adiabatic (iterative solution ret.7) Mixing in the groove 0 0.001 0.002 0.003 0.004 0.005 0.006 in. (£) Michell film shape • Isothermal (exponential solution ref. 5; V-avg12.04 xl0"6lb sec/in.2 + Isothermal (Michell solution ref. 6) pa¥g-- 2.04 x 10~61b sec/in.2 o Isothermal (iterative solution ref.7) P\ - 2.93 x 10"6 lb sec/in.2 x Adiabatic (iterative solution ref. 7 ) Mixing in the groove a Experimental (ret.5) /?= 0.69 radians; /?2 = 11.12 in.; Z//?2 = 0.5618; 8 pads; /V = 3600 rev/min Fig. 9-7. Comparative load-carrying capacity. lations, they give optimistic results and should be used with caution. The adiabatic results agree much closer with practice. This is especially true when mixing in the groove, of the inlet and the hot carry-over oil from the preceding pad, is considered in the analysis. SOURCES 1. Cope, W. E.: The Hydrodynamical Theory of Film Lubrication, Proc. Roy. Soc.} London, A, vol. 197, p. 201, 1949. 2. Cameron, A., and W. L. Wood: Parallel Surface Thrust Bearing, Proc. Sixth Intern. Congr. Appl. Mechanics, 1946.
Adiabatic Solutions 305 3. Osterle, F., A. Charnes, and E. Saibel: On the Solution of Reynolds Equation for Slider Bearing Lubrication. VI: The Parallel Surface Slider-Bearing without Side Leakage, Trans. ASME, vol. 75, p. 1133, 1953. 4. Osterle, F., A. Charnes, and E. Saibel: On the Solution of the Reynolds Equation for Slider Bearing Lubrication. IX, Trans. ASME, vol. 77, pp. 1185-1187, 1955. 5. Sternlicht, B., and H. J. Sneck: A Numerical Solution of Reynolds Equation, Lubrication Eng., vol. 13, no. 8, p. 459, 1957. 6. Michell, A. G. M.: The Lubrication of Plain Surface, Z. Math. Physik, Bd. 50, pp. 123-137, 1904. 7. Sternlicht, B.: Energy and Reynolds Considerations in Thrust Bearing Analysis, Conf. on Lubrication and Wear, London, 1957.
CHAPTER 10 ELASTICITY CONSIDERATIONS INTRODUCTION The preceding chapters deal with completely rigid surfaces and in them it is shown that hydrodynamic pressure generation is dependent on film profile. In Chap. 9 it is shown that, unless variable viscosity is considered in analysis, parallel plates will generate no hydrodynamic force. In practice it is virtually impossible to obtain parallelism; for pressure or temperature gradients will deform the surface. Since the fluid film thickness in conventional bearings is of the order of 10~3 in., the pressure gradients under some conditions can produce elastic defor¬ mations of the same order of magni¬ tude. Thus, for a rigorous analysis of hydrodynamic lubrication, elastic deformation must be considered. ONE-DIMENSIONAL SOLUTIONS 10-1. The Perfectly Elastic Jour¬ nal Bearing. We shall first consider a limiting case in which the bearing is entirely devoid of rigidity, by which is meant that the bearing con¬ sists of an extremely flexible foil stretched around the circumference of the journal such as shown in Fig. 10-1. The film separating the foil from the journal will consist of the following parts: 1. The parallel part, extending over half the circumference of the journal 2. The leading part, at the intake to the parallel part 3. The trailing part, at the outflow of the parallel part The Reynolds equation ^ = G nUh—~ (10-1) dx h* 306 Fig. 10-1. Pressure distribution of foil bearing.
Elasticity Considerations 307 is valid for all three parts of the film. For the parallel film, h — ho and consequently dp/dx = 0; hence, the pressure is constant. For sufficiently low values of ho, say, for ho/R < 10-2, the pressures are generated mainly in the region immediately before the beginning of the parallel film, and the wedge formed between a plane surface and a cylinder may be replaced by a parabolic wedge. Thus Equation (10-1) can be integrated by setting x2/2Rho = tan2 It is then found that where p0 represents the specific pressure, which may be written po = s/R (s = tension in foil). For the purpose of comparing the operational reliability of the foil bearing with that of the conventional bearing, the value of hmin for a rigid bearing with film thickness as given by Eq. (10-2) is where 1/r = l/R — 1/(R + C) « C/R2 (if C/R, the clearance ratio, is sufficiently small). Thus Eq. (10-5) becomes Equations (10-4) and (10-6) enable us to compare the minimum film thickness of the two bearings at high loads. To this end the ratio ho/hmia is determined, and it is found that (10-2) ~ !sin4*)+Ci (io-3) By using the boundary conditions p = 0 for x = oo or \p = t/2 and p = po for x = 0 or ^ = 0 and substituting them into Eq. (10-3), we get (10-4) hmin 2AbLnU r W (10-5) (10-6) It follows that for a clearance ratio C/R of, say, 10-3 the foil bear¬ ing gives greater film thickness than the classical bearing when 10VN/po < 0.024, which is below the acceptable limits of safe bearing operation.
308 Theory of Hydrodynamic Lubrication From this point of view, then, the foil hearing has no advantage over the conventional bearing. Throughout the parallel part of the film the pressure is constant and, consequently, the velocity distribution is linear across the film. For the parallel part the friction exerted on the foil may be written In the other two parts, the velocity distribution is the result of both shear the pressure component drags the foil forward in the trailing part, in the leading part it restrains it with equal force. This component of stress, therefore, does not contribute to the total friction. The friction due to shear in the leading part adds to that in the trailing part, and the sum of their contributions to the total frictional force may be written The summation of Eqs. (10-7) and (10-9) gives the total drag in the foil. The coefficient of friction may be written For yN/p0 < 10-4, the second term represents a correction of less than 4 per cent and may be neglected. One thus sees that in this region of yN/po the coefficient of friction increases less rapidly with yN/p0 for the foil bearing than duced by the deformation of the bearing is small compared to the inclina¬ tion of the slider. The slider-bearing configuration to be investigated is as shown in Fig. 10-2. Fj = 2v*L»NR ^ h o (10-7) and pressure variation along the film. While the friction generated by (10-8) By making use of (10-2) and integrating, it is found that (10-9) (10-10) Springs' ^Bearing Bearing a flexible plate on an elastic foundation. ZI1 / 10-2. Spring-supported Thrust Bearing. ^7 Thrust bearings can be built by mounting for the conventional bearing in which in the region considered, the coefficient of friction is proportional to (yN/p0)**. Under load the plate will deflect, producing a wedge-shaped film. The following deriva¬ tion is for the case in which the wedge pro-
Elasticity Considerations 309 Even the one-dimensional Eq. (10-1) is difficult to solve, since the film thickness is now a function of pressure. A solution will be obtained by starting with the solution for a rigid bearing and then computing a “correction.” This correction will be small, since we are restricting ourselves to the case in which the effect of plate deformation is small. The Rigid-bearing Case. For a slider bearing with the bearing rigid the film-thickness variation is given by ho = ^2 £l “b (a, — 1) —g—j (10-11) The zero subscript will be used to denote the fact that the bearing is being considered rigid. For this case, Eq. (10-1) takes the form t?-O'" («?"») (UM2) where hoo is a constant of integration and is integrated to give OyUB (a - 1\ (x/B)(I - x/B) P" ht* Vo + 1/ [1 + (a - 1)(1 - x/B)r- (1(M3) In this equation the integration constants have been evaluated by the conditions that p = 0 at x = 0 and x = B. From Chap. 3 hoo = hz (10-14) and the load capacity is nr _6nULB2 1 /, ^,<1-1^ /1rt1fN 0 fe22 (o-l)2V o+l/ (10-15) The Flexible-bearing Case. Now, for the slider bearing with a flexible- bearing plate, we will write h = h0 + he, p = p0 + pc, and W = W0 + Wet where he, pc, and We are small quantities which account for the flexibility of the bearing. By substituting these expressions into Eq. (10-1), taking account of the smallness of the corrections, and subtracting off Eq. (10-12), we are left with TT [*•(-* + 3 £)-*■•] (1<M<i> where hco is constant of integration for the correction By substituting Eqs. (10-11) and (10-14) into Eq. (10-16), we obtain dpc _ 6nU 1 di ~ W [1 + (a - 1 )(B - x)/B\* |-2 + 6 ^ p^a irj)-(7j—“j/jj j " (10-17)
310 Theory of Hydrodynamic Lubrication The film-thickness correction hc is the deflection of the bearing plate due to the pressure distribution p0; it can be found by a separate analysis. Equation (10-17) will then yield (upon integration) the pressure correc¬ tion pei and by integration of pe over the slider 1^* ^ ^ length the load-capacity correction Wc can be found. —=—- To determine the deflection of the bearing plate, it be necessary to consider the supporting /o springs as an elastic foundation with a certain „ , modulus k. The configuration of the bearing- Fig. 10-3. Bearing de- . ® flection. plate deflection is shown m Fig. 10-3. To find hCf we must first find y, the deflection of the plate from its original unloaded position. In general, this deflection can be written y(x) = Jf i(x,s)p0(s) ds (10-18) where i(x,s) is an influence function giving the deflection at x due to a unit load at s. This influence function can be shown to be given by i(x,s) = e-^-^cos 0(s — x) -f- sin 0(s — x)] (10-19) with = fc/4D, where D is the flexural rigidity of the plate per unit width given by D — Et*/12(1 — vp2). In this expression, E is the modu¬ lus of elasticity, vp is Poisson’s ratio, and t is the plate thickness. By previous restrictions, y will be taken to be very small. By expand¬ ing the influence function into a power series in 0 and neglecting higher- order terms, we obtain the following simplified expression *(»,*) = ^ [i - 02(s - z)2] (10-20) By substituting Eqs. (10-20) and (10-13) into Eq. (10-18) and integrat¬ ing, we obtain for plate deflection »- HcKm)l|B'" - + 2/3*(B/„ - In){B - x) + /32(/22 - BIn){B - *)*} (10-21) [1 + (a — 1)(B — s)/B]n The deflection y0 shown in Fig. 10-3 is the above expression with x set equal to zero. By subtracting t/0 from y to obtain hCf we have K = lMB ~x) + MB ~x)2] (10-22) where Ai = 2(BI22 — ^32) and A2 = (I22 — Bin).
Elasticity Considerations 311 Thus the film-thickness correction has been determined. It now remains to substitute this into Eq. (10-17) and integrate to obtain the pressure correction. When this is carried out, we obtain 36 v2pU2a2{a 7 r r , Pc kB2hit (a + 1)J /i3Jo3) — CAi (/03J14 — IuJ03) + 2A2(/o3-/23 ~ liiJos) - 6A2 (I03Ju - /2</o3)j (10-23) where the J*s are indefinite integrals defined by -JC (B — x)m , [l + (a- l)(B-x)/B]»ax The new integration constant and hc0 have been evaluated by the bound¬ ary conditions that pe = 0 at x = 0 and x = B. From Eq. (10-13) the dimensionless pressure distribution for a bearing is given by <10•24, 1 - Poh22 , _ B — x where p 0 = —j-g and x — —g— The dimensionless load capacity from Eq. (10-15) reduces to ^•-5Ar.(ln“-2jTi) (ll>-25> .here If the elasticity of the bearing is taken into account, the correction to the film thickness is given by po(s) ln where hc = — X heEh22 ds (10-26) fxUB2 By substituting this into Eq. (10-17), we obtain the form dpc _ a hc{-2 + 6a/(a + 1)[1 + (a - l)x]} - hc0 /in 0>7N dx [1 + (a - l)x]2 ( ° } where v = wnere Vt
312 Theory of Hydrodynamic Lubrication m/B 0.11 § 0.10 ^ 0.09 0.08 r \ / \ / r \ 1.00 1.50 2.00 2.50 3.00 3.50 o Fig. 10-4. Film-thickness correction along Fig. 10-5. Maximum film-thickness cor- the slider. reotion. Fig. 10-6. Pressure distribution in plane Fig. 10-7. Load capacity of clastic slider sliders, (a) Rigid surface; (/>) elastic bearings, (a) Rigid surface; (b) elastic surface correction. surface correction.
Elasticity Considerations 313 By integrating first for the pressure correction and then for load capacity, we obtain WcEhS n2U2LB* (10-28) Note that in this dimensionless form the load capacity and load capacity corrections depend only on a. The above equations are evaluated and plotted in Figs. 10-4 to 10-7. 10-3. Pivoted Shoe with Elastic Defor¬ mation. For the pivoted-shoe bearing shown in Fig. 10-8, where the shoe is con¬ sidered a flexible plate, the shoe deflection correction h. will not be zero, and its value will now be calculated. It will be noted that h, is simply the plate deflection meas- Fig. 10-8. Nomenclature pivoted-shoe bearing. for ured vertically upward and due to po given by the familiar equation (10-29) (Ph. dx2 M D where D is flexural rigidity, as defined previously, and M is the bending moment at x given by M = L /; p0(x — s) ds M Since and p o = L /; p0(x — s) ds — lF0(x — f) £ < x < B fifiUB (a - 1\ (x/B)(I - x/B) V* +V (10-30) f [1 -f- (a — 1)(1 -x/B)\2 1I7 GnULB2 1 / a — l\ M « = —n— 7 I In a — 2 ——r ) h22 (a - 1)2\ a + 1/ the expression for flexural rigidity can be integrated directly for h. and written in the functional form h. = nUB* I)h22 (10-31) where, in order to compare the rigid-slider case with the flexible shoe case on the basis of the same minimum film thickness, the deflection of the slider at a; = B is set equal to zero. One integration constant Ci remains; it is determined by the condition that the center of pressure remain at x = £ when the influence of the flexibility of the shoe and runner on the pressure distribution is taken into account.
314 Theory of Hydrodynamic Lubrication For the configuration shown in Fig. 10-8 where both the runner and the shoe are flexible the film-thickness correction hCf given by the deflection correction of the runner hT (which here corresponds to the bearing deflec¬ tion analyzed in the previous section) plus the shoe h9, can be written in the form nUB2\ he = Eh2 where [f(a’£) + %g(a’Ci’i)} _ (il/B)> (10-32) 1 In order to find the correction pe to the lubricant pressure p0, this value of he must be substituted into Eq. (10-16) and the resulting equation Fig. 10-9. Film-thickness correction of elastic pivoted-shoe slider for x = 0.5687. Fig. 10-10. Pressure correction for elastic pivoted-shoe slider for x = 0.5687. integrated. The constant Ci is determined by requiring that the center of pressure remain at x = £. The load capacity correction We is then given by Eq. (10-28). This problem was solved on a digital computer. The results of the computations are shown in dimensionless form in Figs. 10-9 and 10-10. In Fig. 10-9, the film-thickness correction hc — hr h8
Elasticity Considerations 315 is plotted in dimensionless form against distance along the slider x/B with a as a parameter for an £ value of 0.5687 (a = 2). In Fig. 10-10, the pressure correction pe is plotted in dimensionless form against dis¬ tance along the slider x/B with a as a parameter for an £ value of 0.5687. For this value of £, the load capacity correction Wc is given as a function of a in the following table. Note that for a sufficiently small a the shoe Load Capacity Correction as a Function of WeEh2> n*U*B*L 0.1 0.475 0.2 0.0465 0.3 0.00163 deformation counteracts the runner deformation to the extent that the net effect is an increased load capacity. As the shoe becomes thicker (a increases), the load capacity increase becomes smaller. For an infinitely thick shoe, the load capacity correction is equal to that for a rigid slider, which for this value of £ is given by W.- -0.00903 TWO-DIMENSIONAL SOLUTIONS OF CENTRALLY PIVOTED SECTORS A centrally pivoted sector thrust bearing can carry load and maintain equilibrium of moments by virtue of either of two processes: a thermal wedge or elastic deformation. Hydrodynamic pressures can also be generated and equilibrium of moments satisfied in a pivoted-sector thrust bearing by offsetting the pivot. This design yields optimum character¬ istics; however, it can be used only in applications where one direction of motion is present. The table below gives comparative results for a 31-in.-OD 153^-in. ID bearing with a subtended angle of 38J4° operating at 320 rpm with a minimum film thickness of 0.001 in. and an inlet oil viscosity of 5.1 X 10-6 lb-sec/in.2 at 130°F. (At 210°F this oil has a vis¬ cosity of 12.3 X 10~6 lb-sec/in.2) It is seen that the thermal-wedge effect Load, lb Maximum temperature, °F Thermal wedge (flat pad, central pivot) 22,200 204 Elastic deformation (central pivot) 36,600 185 Optimum pivot (0 = 0//3 =61 per cent) 39,600 179 allows the flat centrally pivoted pad to carry approximately 57 per cent of the load carried by the pad with optimum pivot. At the same time,
316 Theory of Hydrodynamic Lubrication the maximum temperature reached with the flat centrally pivoted pads is 25°F higher than that reached in the pad with optimum pivot. How¬ ever, experience suggests that the difference in performance between central and optimum pivot location is not so severe. By considering a simplified elasticity solution which allows pad deformation to be approxi¬ mated and included in the analysis, the results between theory and Fig. 10-11. Coordinate system of sector thrust bearing with radial and tangential deformation. 170 u_ °«r 160 3 I 150 o. ? 140 c 1 130 a» | 120 o 110, —- y Point! ;reter to test 1 result s cone lucted — by three different investigators Ret.5 Tr- 115°F, fi at 100°F = 12.5x10'6lb sec/in.2 ft at 210°F = 1.15 x 10"6lb sec/in.2 0 100 200 300 400 500 600 700 800 900 1,000 1,100 1,200 Unit loading, psi Fig. 10-12. Groove mixing temperature vs. unit loading. practice compare more realistically. By comparing the elastically deformed pad, we see that it is capable of carrying approximately 92 per cent of the load and that its maximum temperature is only 6°F higher than that of the flat pad with optimum pivot. The analysis presented here employs numerical methods for the simul¬ taneous solution of the Reynolds, energy, and elasticity equations. In Chap. 9 the simultaneous solution of the first two equations was discussed; thus, here only the added factors will be considered. We
Elasticity Considerations 317 200 400 600 800 IjOOO V200 1,400 Unit loading, psi Fig. 10-13. Chart of maximum and average temperatures; oil viscosity and pad inlet temperature per Fig. 10-12. Fig. 10-14. Chart of load-carrying capacity, G pads, hmin = 0.0010 in.; oil viscosity and pad inlet temperature per Fig. 10-12.
318 Theory of Hydrodynamic Lubrication assume that the convex shape into which the pad bends under load may be represented by part of a spherical surface whose radius of curvature is Ref as shown in Fig. 10-11. In all cases considered here, the pad curva¬ ture Re is very large. In addition, the pad inclines, so that its tangent plane directly above the pivot point has slopes me (circumferentially) Unit loading, psi Fig. 10-15. Chart of load-carrying capacity, 6 pads, hmiu = 0.0008 in.; oil viscosity and pad inlet temperature per Fig. 10-12. and mr (radially), with respect to the plane of the runner, as shown in Fig. 10-11. The pad inclinations are small, so that sin me — tan me — me sin mr = tan mr — mr w cos me = 1 cos mr = 1 In accordance with plate theory, the bending deflections are taken to be proportional to load and inversely proportional to the pad thickness cubed.
Elasticity Considerations 319 Fig. 10-16. Chart of load-carrying capacity, 6 pads, hmin = 0.0006 in.; oil viscosity and pad inlet temperature per Fig. 10-12. Let the film thickness at a reference point (xa,ya) on the pad surface be ha. The film thickness at any other point (x,y) can then be written h = ha - mt(x - x„) + mr(y - y«) + Rc ^1 - X° ^2Va ^ ' -(‘-ttT] (,0-34) Since Re is very large, powers of the ratio r2/Rc2 are neglected. Equation (10-34) then becomes h = ha — me(x - x.) + m,(y - ya) + {x* + y,) + y‘] (10-35) This equation can be converted from the x,y coordinate system to the r,d coordinate system of Fig. 10-11 by means of the relations
(10-36) 320 Theory of Hydrodynamic Lubrication y = r cos ^0 - 0 - f cos ^0 - ^ The general equation for the film shape in polar coordinates is then h = ha + m91 ra sin ^0O — ^ - r sin ^0 — 0 j - mr | r0 cos ^0a - ^ — r cos ^0 — 0 j -f- [r2 — ra2 — 2rr cos (0 — 0) + 2rar cos (0a - 0] (10-37) Equation (10-37) can also be used to describe the film shape for flat pads. In such cases, Rc is infinite, thus eliminating the fourth term on the right-hand side of the equation. a4S XX X X X xxxxx xxxxxxxx Unit loading, psi Fig. 10-17. Chart of load-carrying capacity, 8 pads, hmin = 0.0010 in.; oil viscosity and pad inlet temperature per Fig. 10-12.
Elasticity Considerations 321 Unit loading, psi Fig. 10-18. Chart of load-carrying capacity, 8 pads, hmin = 0.0008 in.; oil viscosity and pad inlet temperature per Fig. 10-12. For cases in which loads are light and the bending deflections are small, it is convenient to use as reference the point at the inside radius and trail¬ ing edge of the pad. Equation (10-37) then becomes h = hi + me (R — L) sin ^ — r sin ^0 — ^ j — mr[(R — L) cosf - r cos (<> - §)] + ~ |V — (R — L)2 + 2(11 - L)r cos ~ - 2rr cos (e - (10-38) For cases in which the loads are large and the bending deflections are of the order of the minimum film thickness, the point of minimum film thickness may fall within the pad boundary. It is then more convenient to use this point as reference. Equation (10-37) then becomes
322 Theory of Hydrodynamic Lubrication h = ^ j + r [«, sin («-§) + (ic ~ *,) cos («-§)- ggj (10-39) and the coordinates of the point of minimum thickness are Tm = Rc j m,2 + j 0 , z . Ttl) 6m - 2 + tail f/R' _ ^ (10-40) Equation (10-35) and Fig. 10-11 show that (with the simplified elasticity approach used here), the bending deflection at any point on the pad surface is proportional to the square of its distance from the pivot, that is 1 5 = 2WC + (10-41) Unit loading, psi Fig. 10-19. Chart of load-carrying capacity, 8 pads, hmia = 0.0006 in.; oil viscosity and pad inlet temperature per Fig. 10-12.
Elasticity Considerations 323 Unit loading, psi Fig. 10-20. Chart of load-carrying capacity, 10 pads, hmln = 0.0010 in.; oil viscosity and pad inlet temperature per Fig. 10-12. The value of the bending coefficient 1/2Re was obtained by calculating the deflection at the rim of an equivalent circular plate, point-supported at the center of its lower face and carrying a conically distributed load on its upper face. A circular plate was used because a closed solution for its bending deflections is available.8-* A conical load distribution was selected because sinusoidal and parabolic pressure distributions yield essentially similar results and because the ratio of peak to average pres¬ sure (3:1) is similar to that in an actual bearing pad. By integrating Eq. (57) of Ref. 8 for a steel circular plate (radius a and thickness ^vg) under the loading and support described above, the deflection at the rim is found to be Wa2 8 = 0.225 (10-42) * Such superscript figures indicate references listed under Sources at the end of the chapter.
324 Theory of Hydrodynamic Lubrication Unit loading, psi Fig. 10-21. Chart of load-carrying capacity, 10 pads, hmin = 0.0008 in.; oil viscosity and pad inlet temperature per Fig. 10-12. From Eqs. (10-41) and (10-42), the relation between the bending coeffi¬ cient and the pad load is 1 W 2 Rc 0 225 tl'E (10-43) At each operating point, the pad deformation has to be related to the pad load in accordance with Eq. (10-43). The film shape which depends on this deformation and on the inclinations of the pad has to be such that the resulting center of pressure passes through the pivot. Finally, the inlet boundary temperature has to be related to the unit loading in accordance with Fig. 10-12. In order to meet these requirements, the following trial-and-error procedure has to be used: 1. For the bearing geometry being studied, select a value of minimum film thickness. 2. Estimate the corresponding unit load and hence the inlet tempera¬ ture Ti and the bending coefficient 1/2Rc.
Elasticity Considerations 325 3. Select values of radial and tangential inclinations (me and mr, respectively). 4. Introduce the above as input data and obtain the corresponding solution. 5. Determine the coordinates of the center of pressure and the actual unit load (and hence the actual inlet temperature and bending coefficient). Check whether these agree with the estimated values within the following prescribed error limits. In the results presented later, the following accuracies were imposed: a ITw, - ru < 2°F b c !(?%)«,... - f%I < 0.5% d |(0%)«.„.i - S%\ < 0.5% (iL... - GklJ ^2 x ,o-6 in - (io-44) "•s ^ ^ in 5 s oo Unit loading, psi Fig. 10-22. Chart of load-carrying capacity, 10 pads, and pad inlot temperature per Fig. 10-12. hmin = 0.000(> in.; oil viscosity
326 Theory of Hydrodynamic Lubrication If any of the conditions a to d of Eq. (10-44) are not satisfied, steps 2 to 5 must be repeated until all errors are within the specified limits. Figures 10-13 to 10-24 give computer results for bearing sizes ranging from 19 to 50 in. in diameter. These bearings are all geometrically similar, with the following properties: £- = 0.5 It 2 ^ = 0.154 f% = 6% = 50 "D W U VJ.icrx , /0 — V /0 — w It 2 It 2 with 15 per cent of the bearing surface taken up by oil inlet grooves. 20 10 r 5 D ; 2 1 0.5 0.2 6 pod (0=51°) 8 pad IP = 38.25°) 10 pad (P = 30.6°) \ UJ I I I I I 1 I 0.02 0.05 0.1 0.2 0.5 1 - /*ovg ^ovg PB X 106 Fig. 10-23. Chart of hydrodynamic oil flow per pad. Average film temperature per Fig. 10-13. Fig. 10-24. Chart of horsepower loss per pad. Average film temperatures per Fig. 10-13. When the oil-film temperatures are plotted, it is seen that both the maximum and average temperatures are, with good accuracy, functions only of the unit load, number of shoes, and pad inlet temperature. This allows the maximum and average temperature to be represented on a single chart (Fig. 10-13). The accuracy of this chart, up to T— = 235°F, is ±5°F. Above T^ = 235°F, the accuracy is ± 10°F. Numerical methods for evaluation of pressure gradients are not too accurate. Thus errors as high as 20 per cent may exist in flow and horse¬ power results plotted in Figs. 10-23 and 10-24, respectively. SOURCES 1. Blok, H., and J. J. van Rosaus: The “Foil Bearing”: A New Departure in Hydro- dynamic Lubrication, Delft Publication 140, Dec. 18, 1952. 2. Osterle, F., and E. Saibel: The Spring Supported Thrust Bearing, Trans. AS ME, vol. 79, February, 1957. 3. Osterle, F., and E. Saibel: The Effect of Bearing Deformation in Slider Bearing Lubrication, Trans. ASLE, vol. 1, no. 1, 1958.
Elasticity Considerations 327 4. Osterle, F., and E. Saibel: Surface Deformation in Hydrodynamic Slider Bearing Problems and Their Effect on Pressure Development, Conf. on Lubrication and Wear, Paper 35, London, 1952. 5. Sternlicht, B., and E. Arwas: Propeller Shaft Thrust Bearing Analysis, Phase I, Bur. of Ships, U.S. Navy, 1959, TIS 59 GL 81, General Electric Company. 6. Sternlicht, B., J. C. Reid, and E. B. Arwas: Performance of Elastic, Centrally Pivoted, Thrust-bearing Pads, Part I, ASME Paper 60-LUB-10. 7. Sternlicht, B., G. K. Carter, and E. B. Arwas: Adiabatic Analysis of Elastic, Centrally Pivoted, Sector, Thrust-bearing Pads, ASME Paper 60-WA-104. 8. Timoshenko, S., and S. Woinowsky-Krieger: “Theory of Plates and Shells,” 2d ed., McGraw-Hill Book Co., Inc., New York, 1940.
CHAPTER 11 HYDRODYNAMICS OF ROLLING ELEMENTS GENERAL REMARKS One of the more complex problems in the field of lubrication is that of gear teeth and rolling-element bearings. This problem involves not only the theories of elasticity and hydrodynamics but also the consideration of such effects as heat transfer, relaxation time, compressibility, and the variation of viscosity with pressure, temperature, and rate of shear. Knowledge in some of these areas is very meager, particularly as regards the effect of shear on viscosity under high pressure and the relaxation time of lubricants. The subject of lubricating rolling elements will be discussed beginning with very simple cases and progressively increasing the complexity of the discussion as the effects of the individual factors are brought in. Thus we shall consider first a case which deals entirely with hydrodynamic theory and neglects heat transfer, compressibility, and elastic effects. The theory of lubrication as developed by Osborne Reynolds for bearings appears to be equally applicable to the lubrication of high¬ speed gears and rolling-element bearings. The absence of wear must be attributed to the presence of a fluid film between the meshing surfaces. Because the contact zone between the teeth of a gear is very narrow, there is no appreciable fluid flow parallel to the line of contact; the entire flow takes place in the direction of motion. Thus we can apply one-dimensional analysis to the treatment of gears, whereas in considering bearings a two-dimensional analysis is more applicable. In the case of the usual type of teeth found in gears, the motion of the teeth relative to each other is primarily that of rolling, the slip being only a fraction of the roll. A simple modification of the conventional solu¬ tions is applicable to this case. Suppose that in Fig. 11-1 there is a roller of radius R revolving with a surface speed U, and that at a short distance below the roller there is a plane surface moving with the same speed. If the roller and plane are covered with fluid, they will drag 328 Fig. 11-1. Hydrodynamics of cylinder and plane surface.
Hydrodynamics of Rolling Elements 329 this fluid along, crowding it into the constantly narrowing space. The result is that a pressure is developed in the fluid. As will be demonstrated later, this pressure may attain extremely high values, although it is confined to a very narrow band. FLUID FILM WITH RIGID SURFACES 11-1. Solutions with Constant Viscosity. For conditions of pure roll¬ ing found at the gear pitch line, the theory is the same as for a roller bearing without slip. Since end leakage is neglected, the equation for the pressure gradient is, from Chap. 1, g = 6 (11-1) The film of a roller above a flat plate may be expressed by (2R — h + hmiB)(h - hmin) = x2 If we neglect the square of h — hmin, which is very small in the neighbor¬ hood of the point of contact, the above equation gives h = A"‘" (* + 2Rh~) (ll'2) Substituting Eq. (11-2) in Eq. (11-1) and integrating, we get p = 3m(£/i + U2) sec2 *°) + * + sin * cos — sec2 Xo sin x cos3 xj where Xo corresponds to h0. The above equation can also be written in the following form p = UV2Rh~\ x sin 2x tr 12m 1.2 ~4 4 r (‘ix 3ir sin 2x sin 4x\l ClV8-l6 + -ir +^i2~)\ (11'3) where Ci is an integration constant, Ui = L\ = U, and tan x = x/y/2Rhmin To determine Ci, we note that, when the pinion tooth begins to engage the rack tooth, the pressure is zero when x is zero. From the pressure generation standpoint this is an unfavorable condition as compared to the case represented by Fig. 11-1, where the region of pressure extends also
A 0 330 Theory of Hydrodynamic Lubrication to the left of OE. Taking it as one possible case, however, we note that, when x = 0, x is also zero, so that putting p = 0 in Eq. (11-3) with this value for x gives Ci = %. We then get V U (20. , iy . ^ U(2Rhmia)»<t> ih'-inL 12 (sm2x + ^sin4x)“ —35:— The minus sign appears because, as shown in Fig. 11-1, U is directed toward negative x. The above function is evaluated in Table 11-1. If % > Ci > 1, we obtain a curve of the general type shown in Fig. 11-2. Here the curve cuts the x axis to the left of the origin, at point B an<£ again at A. Frofilefor max^mum pressures are attained when points B r and A coincide. The curve then touches the x axis without cutting it. The distance of this point of contact from O is given by the relation tan x = —r—— = -0.47517 V2Rhm>. and the value of Ci is 1.2258. If we denote this curve by \f/, then the pressure at any point is given by V = ¥^V2Rh~t Values of are given in the fourth column of Table 11-1. If we take C\ as unity, we obtain 1 U \Z2Rhmia (it _ , sin 4x 12m P A*,. 8 (7r _ , sin 4x\ 2-‘ + —) The curve thus expressed has a point of inflexion when x = 0. While this curve does not correspond to any practical lubrication case, if we denote 1 Ar _ sin 4x\ . 8 \2 _ X 4 / = 8 then every possible pressure curve (subject to the condition that the pressure is zero at infinity) may be expressed as j= ~ ^2Rh~ [(1 - y)6 + y<t>] For the curve which we have denoted by \f/, y has the value 0.67725. Values of 0 are given in the third column of Table 11-1.
Hydrodynamics of Rolling Elements Table 11-1. Pressure Distribution Functions a -si a > -0.5 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.5 2.0 2.5 3 4 5 6 7 8 9 10 H2 ii •O- -0.10667 -0.09908 -0.08417 -0.06163 -0.03266 0.00000 0.03266 0.06163 0.08417 0.09908 0.10667 0.10813 0.10510 0.09915 0.09158 0.08333 0.04734 0.02667 0.01585 0.01000 0.00589 0.00246 0.00146 0.00093 0.00063 0.00045 0.00028 •H .5 *35 + I k l<N <-< 100 II 0.22431 0.21270 0.20406 0.19883 0.19668 0.19635 0.19602 0.19387 0.18864 0.18000 0.16839 0.15474 0.14011 0.12540 0.11129 0.09818 0.05132 0.02796 0.01635 0.01022 0.00467 0.00223 0.00147 0.00094 0.00063 0.00045 0.00033 •B- LO CO a ■« S > d v JG O 0.000155 0.00154 0.00885 0.02244 0.04135 0.06337 0.08538 0.10430 0.11788 0.12519 0.12658 0.12317 0.11639 0.10762 0.09794 0.08812 0.04862 0.02708 0.01601 0.01007 0.00549 0.00239 0.00146 0.00936 0.00632 0.00445 0.00029
332 Theory of Hydrodynamic Lubrication The total area of 0 between x = 0 and x = + <» is easily obtained; it is equal to y/2Rhmin/4i. Similarly, the integral of <f> between the same limits is y/2Rhmiu/Qf while the value of ^ between x = — y/2Rhmitk X 0.47517 and x = oo is 0.2040 y/2Rhmin, or a little over y/2Rhmtn/5. With this latter curve the load carried with a given thickness of oil film is a maxi¬ mum. This total load is L [' pdx = - X2^~L V2Rh~ y/2Rh~. X 0.2040 J-x nmxa = -4.896 ^RL (11-4) The maximum pressure is attained when .r„ = 0.47517 y/2Rh~, and its value is Po = 1.521 -rj— ^2Rh~ ^min These expressions apply to a pinion gearing with a rack. If, however, the pinion meshes with a gear, we must replace R by p, where I = _L+_L p R\ Ri In this expression R\ denotes the radius of curvature of the pinion tooth and Ri denotes that of the gear tooth. The work done on both acting surfaces is dH = 2UnL ^ I dx dy |o 12I/VL r 1 Cx dx hm, „ [1 + xy2Rh„in (1 + x*/2Rh„t.) which when integrated yields (11-5) The following boundary conditions such as may be found in a cam and cam follower are next considered u(x, 0) = l't u(x,h) = L'i v(x,0) = Vt v(x,h) = V, The Reynolds equation (l-6a) without side leakage becomes 1 d/h°dp\_ d - f/2) I ,r i2d^\Jdl) “ ~dJ-[ J + '1_ u
Hydrodynamics of Rolling Elements 333 The cam profile can be approximated by a parabola. The oil-film thickness is then expressed by Eq. (11-2). If the follower does not move in the xz plane, and if the surface velocity of the cam is U, then U2 = 0 Ui = U = 0 For these boundary values, the above equation can be integrated to give ,f,Ln ■3' _ 2) + »v2B)*^+1“ (se: + sir.) [fc, + .V2R + (£)“ "c“" vfa] + in which C1 and C2 are unknown constants of integration. The value of hmin is also unknown. These quantities will next be evaluated. If the velocity U is positive, the cam and oil are moving across the follower from the negative x direction to the positive x direction. Thus the oil in the negative x direction has not yet passed between the cam and follower, and, if sufficiently far away, the oil is at atmospheric pressure. Consequently, lim p(x) = 0 X—► — 00 For the conditions to be met, the constant C2 must be and Eq. (ll-6a) becomes PM - 3, (j-1 - + .i/jsji + !' (ar + 4l—) +V/2* + (£)“ (1+ vit)] <n-IIW As the oil is carried by the cam through the narrow space between the cam and follower, the pressure rises to a maximum and then decreases. Eventually, after the oil passes the region of minimum film thickness, the pressure drops to zero, but cavitation prevents it from becoming negative. Consequently, for some value xe which corresponds to the end of pressure zone, and the pressure remains zero for all larger values of x. These boundary conditions cannot be used to evaluate the remaining unknown constants
334 Theory of Hydrodynamic Lubrication Ci and /tmin explicitly. However, Ci can be replaced by a function of the unknown xe, since dp _ 3yUx2/R + 12/zCi . .. dx~ (hmiu + x*/2Ry KLL"n must vanish for x — xe\ that is, n - Ux'~ 1 4 R Equation (11-66) then becomes _ 3yUx (hmln + x.*/2R) 3yU ( 3x«2 \ [ * 26mln (6min + z2/2fl)2 ^ 4/imin V 2hmiaR) + x2/2R + (E)"(i + “ct“vfa)] The pressure p(x) given by Eq. (11-8) meets the condition that |dp/dx|*# is zero. If a new variable xa is defined by - Xe Xa — V2 hml.R the additional condition that p(xe) itself vanishes can be reduced to — xa — 3xaz + (1 — 3xa2)(l + xa2) ^ + arctan xa^ = 0 The only real root is Xa = - j-*'— = 0.475130 (11-9) V2haiaR Both xe and haln are unknown, but if either one is somehow evaluated and the other one is determined from it by Eq. (11-9), then p(xe) is zero. Another condition can be imposed to determine hmia: the pressure must be able to support the force W between the cam and follower. Thus, if L is the length of the contact between the cam and follower, then W = L p" p(x) dx in which p(x) is given by Eq. (11-8). Upon integration, j1^1 ~20) [l + (2 + arctan V2^r)]} Equation (11-9) can be used to evaluate xe/\/2hmiaR. Then ij/ _ 3yURL ~2hZT
Hydrodynamics of Rolling Elements 335 Equation (11-7) implies that +xe and —xe are the only finite zeros of dp/dx. Since dp/dx is negative between x« and —xe and positive else¬ where, p(x) is a maximum for whichever value of +xe and — xe is nega¬ tive, and p{x) is a minimum for whichever value is positive. The maximum is p(—xe) because Eq. (11-9) shows that +xe is positive. From Eq. (11-8) / \ = 3fxUXe 1 3pU [* _ 3xe2 1 [ Xe Xe) 2hmin hmiD + x*/2 R 4/tmin L • 2hm^R\ |>min + x*/2 R (r—)^ (arctan — -)1 \hminj \ \/2hminR 2/J Equations (11-8) and (11-10) can be used to eliminate xe and hmin: WH Po = p( X*) = 0.280879 (11~11) 11-2. Viscosity as a Function of Pressure. If the variation of viscosity with pressure is represented by the form p = poeap, then UM2, The right-hand side of Eq. (11-3) will be the same; the left-hand side will be 1 — e~ap/a instead of p -*P = 3(t/i + C/2)Mo (11-13) ® “'min (1 + % sec2 xo) + x + sin x cos ^ H sec2 x0 sin x cos3 x j and 1 — e~apo = 0.76mo«(C/i + t/2) We see that po will be infinite when the right-hand side goes to unity. The total load W is again J^Lpdx, and when po~* <*>, W is still finite though about two and a half times the value corresponding to the p = po case. If the speed is increased, the maximum pressure cannot rise (because it is already infinite), but the peak, instead of being infinite over an infinitely small area, becomes infinite over a finite area. Thus the film thickness hmla must increase (assuming a and mo are constant). Hence the disks become oil-borne whatever the load. This accounts for the observation that the friction-speed curve is independent of load; i.e., the drop from mixed friction to full fluid friction is independent of load and dependent only on speed. The obvious criticism of this analysis is that an infinite pressure will produce a number of other phenomena neglected in this simple treatment:
336 Theory of Hydrodynamic Lubrication an infinite pressure will cause infinite viscosity, and at that point infinite heat generation, which will reduce mo and a; and the disks will deform and alter the shape of the oil film, which in turn will change the pressure curve. FLUID FILM WITH ELASTIC DEFORMATION We next combine the hydrodynamic theory with the theory of elastic¬ ity for two conditions: constant viscosity mo and a viscosity given by m = Moeap. When the disks are not deformed, the film thickness is h = he + 2P where subscript e corresponds to the end of the lubricant pressure zone. As a result of deformation, the displacements Si and s2 at the boundary points of the disks must be added. Since the pressure is important only adjacent to the narrowest point of the gap, we replace the disks by half planes in the calculation of Si or s2. In the case of a half plane4* the displacements s of the boundary points for a given perpendicular load W are determined by the equation E - - «« - »(2' - »i rh * (1 M4> If we select a pressure distribution p as a function of x, we can then determine ds/dx and, from it, by further integration, s. The change of the gap becomes then s9 = Si + s2 = 2s. The resulting force due to pressure p for the disk length L becomes W By employing the dimensionless parameters p = p(L.re/!F), x = x/xe, we obtain p as a function of x: We also define Xo/xe as the point at which po is to occur. Then, Eq. (11-14) becomes (11-15) * Such superscript figures indicate references listed under Sources at the end of the chapter.
Hydrodynamics of Rolling Elements 337 and E 4(1 - O W f* I 1 J, = / - - Jl \ IT By means of the two integrations, we obtain s as a function of x; the second integration constant is so selected that x = 1, also for the end of the lubricant pressure se = 0. The film thickness with deformation then becomes T^ — T ^ h = he + s, H »—- 2p which in dimensionless form may be written as £ = £. + *+ A(x2 - 1) (11-17) E L L , _ E L xe2 where h 4^_Vt>2)wh A 4(1 — v2) W 2p The equation for pressure in the lubricating film is dp = 6 lx(Ul + At the point x = — <» and x = xe, p is zero. At the point x0, where h = ho = hej p = po. The hydrodynamic equation for the dimensionless quantities reads With h according to Eq. (11-17) we have B fP — dp = fZ + dx = /(x) (11-19) > M ./—[/l, + s + X(x*- 1)|3 The displacement function s is therefore determined. The values A and he have to be chosen. The value A is so selected that dp at the pre¬ determined point Xo/xe becomes zero. From this it follows that for this point
338 Theory of Hydrodynamic Lubrication Furthermore, the value of he is so selected that p at the point x = 1 becomes zero or becomes equal to what it is for x = — oo. From this the following condition results: /. s -f A(x2 — 1) -oo [he + S + A(x2 - 1)]: dx = 0 (11-21) Equation (11-19) is evaluated with these two values, and f(x) is deter¬ mined. If the viscosity is constant, that is, y = mo, the left-hand side of Eq. (11-19) furnishes the value Bp] otherwise, mo/m is to be expressed as a function of p and we obtain instead of Bp a function of p, from which p has to be determined. We then compare the final function of p with the initial function of p. If they do not agree, the final function of p is selected as the new initial function. Details of the method are indicated in the following two examples. 11-3. Solution with Constant Viscosity. We select x0/xe = —0.6. For disks without deformation xQ/xe becomes = — 1; with deformation, the point of maximum pressure moves closer to zero. We select p as a function of x and make sure that pQ occurs at the point x = —0.6 and that f^goPdx = 1- Figure 11-3 shows the selected characteristics. From it, the function s is determined according to Eq. (11-16), which is illustrated in Fig. 11-4. We read from this diagram for x0/xe = —0.6 the value s|_0.e = 0.363 and obtain from it, according to Eq. (11-20):
Hydrodynamics of Rolling Elements 339 In Fig. 11-4 is also plotted —A(x2 — 1), so that 5 + A(x2 — 1) can be read off. After several trials, it is found that, according to Eq. (11-21) he = 1.810 Now the integral of the right-hand side of Eq. (11-19) is evaluated. This integral furnishes the function Bp. If we integrate this function from x = — oo to x = 1, we obtain jl_m Bpdx = B = 0.2779 We divide the function Bp by this number and obtain the final function p, which should be the same as the initial function. The trial-and-error procedure is followed until the initial and final functions are equal. 11-4. Viscosity as a Function of Pressure. The beginning of the analysis is the same as for constant viscosity. We again select ^ = -0.6 Xe and assume an initial function for p (Fig. 11-5). From this, we obtain 5 and from it B^dp = B (exp -^pyp = lL^rdi According to Eq. (11-19), we have B j* (exp - g p) dp = f(x) (11-22) The function f(x) is fixed for the selected initial function p.
340 Theory of Hydrodynamic Lubrication Equation (11-22) becomes, after integration of the left-hand side, B^w[1~exp{-iS'p)] =m (11-23) For further analysis we must now determine an additional dimensionless constant. From Eq. (11-23) it follows that - / aW \ aW 1 -exp {- u.p) = bu.s{x) The expression on the left-hand side has the minimum value of 0 for p = 0, and can assume the maximum value of 1 when (aW/Lxe)p becomes Fio. 11-5. Pressure distribution in contact zone for n = p0ca*», n = 0.8. infinite. Normally, however, p will not reach such high values, so that the left-hand side assumes a maximum value of n < 1. Thus t / aW _\ aW f ... n = 1 " exp (- IF.Po) = blfMx) therefore a\V _ n BLXe /mix and Eq. (11-23) reduces to -.ffi!] (,,.24) The function on the right-hand side is determinate, since f(x) and there¬ fore also /mfcX was determined and n was selected.
Hydrodynamics of Rolling Elements 341 From Eq. (11-24) we find «- - -/'.%[i -.£■;]«<*> We then divide the function on the right-hand side of Eq. (11-24) by B and so obtain the final function p. This is compared with the initial function. If necessary, the calculation is repeated with the final func¬ tion as the new initial function. Results are given for n = 0(/x = mo), and n = 0.8. In addition, a calculation for disks without deformation (E = °o),x0/xe = — 1 was carried out separately for arbitrary values of n. Figure 11-3 shows the Hertzian pressure distribution and also the function p(x) for xQ/xe from 0 to — 1 for constant viscosity. The maxima of the curves are connected by a line. For n = 0.8 (Fig. 11-5) the maxima are located somewhat higher because of the effect of pres¬ sure on viscosity. The shape of the film thickness for n — 0.8 is indi¬ cated in Fig. 11-6; for n = 0, the curve is only very little different. In the solution of a physical problem, one must determine what values of xo/xe and n apply. The above analysis gave us the following character¬ istic quantities 4 £ kil 4(1 - >v2) W 2p d = r 4(1 - »p*)i2 (w\* i i [ E \ \LJ 6Mo(t/. + n [ E VfW /-., " Ua-V)JU x, may be eliminated by taking y/A B ^/tb = r4(1 - ^ 1 _ it* [ E \LJ 6m.(I/i+ For a given load and constant viscosity the maximum pressure under hydrodynamic conditions for undistorted disks can be related to the maximum Hertzian pressure by the following relation: 1 1 6/Xo(f/l "f* Ui) Xe2 (11-25) Fig. 11-6. Film thickness in the contact Z\ — \/AB — 1.218p0hydro/p<)Hert. (11-26)
342 Theory of Hydrodynamic Lubrication The next quantity must be so constituted that a occurs in it. In order that the quantity also apply to nondeformed disks, we must combine and B in such a way that neither xe nor E/4(1 — vp2) occurs. The quantity is Z2 = A a/B5 = £ ViW/DWMUt + u,)\ (11-27) J max "P Figure 11-7 shows the curves for n = 0, 0.2, 0.4, 0.6, and 0.8 and the curves x0/xe — —1.0, —0.8 to —0.2, and 0 as functions of Zi and Z2. We can now calculate Zj and Z2 and from Fig. 11-7 determine n and xQ/xe. After n and xQ/xe are found, we can determine the quantity xe. For this purpose, we can use the values for A or those for B We select, however, a quantity which does not contain the modulus of elasticity, i.e., g VCW7L)[l/6w(tfi + Ut)] = A VB (11-28) The values for A y/B are plotted in Fig. 11-8 for n = 0 and 0.8 as a function of xQ/xe. For x0/xe = —1, the relation to n is indicated in the side view. Another important quantity is hmia. For disks without deformation and for constant viscosity it is —TOi? — ?: = o 204 2p L 6Mo(t/i + Ut) A corresponding quantity will be derived for disks with deformation; the equation for this is
Hydrodynamics of Rolling Elements 343 This quantity is plotted for n = 0 and for n = 0.8 in Fig. 11-9, and in the side view on the left-hand side the function n for disks without deforma¬ tion, that is, xo/xe = — 1 is also given. As a final quantity, we shall discuss the maximum pressure; for constant viscosity and without deformation, it is p«2p + Ut) = 1.373 We know from the developed pressure distribution curves the numerical values for poLxJW = po] by combining this with Eq. (11-28), we obtain Po2p + Ui) = Po (11-30) The resulting values are plotted in Fig. 11-10 for n = 0 and n = 0.8. and on the left-hand side for x0/xe = — 1. It will be seen from J = 6M(/i - Uo) (11-31) that if po is constant in the loaded area, that is, dpo/dx = 0, then h is constant and equal to ho. With constant film thickness over the major An in W 1 Fig. 11-9. Diagram of hmia vs. x0/xe and n. Fig. 11-10. Diagram of p0 vs. xq/xc and n. part of the contact area the distribution of normal stresses between the rubbing surfaces will be the same as the semiellipse of normal pressures for elastic contact in the absence of lubricant. We shall now examine the boundary conditions and their effects on lubrication. It is convenient to begin this examination at the entrance, since the processes which take place there determine the value of ho. The pressure curve can be a semiellipse (or any other shape) for a con¬ stant value of h. This will occur only after the quantity p0 In x has reached a value very near unity. Therefore the pressure developed between the diverging parts of the rubbing surfaces before the contact area must be such that the value of pQ ln x is very close to unity at the start of the plane part of contact. From now on the value of p0 ln x is considered to be unity at the start of the contact area.
344 Theory of Hydrodynamic Lubrication According to Ref. 8, with dry friction the shape of the gap between rubbing surfaces outside the contact area is given by the following expression, which is applicable for the initial contact of two surfaces at the point. •'(*) = -%W(Vl + vt) fe (i _ .__f! l4l \ _ _ d<t> _ (11-3*2') Jo \ l>2 + <(> a2 + 4>)y/(at+ *)(6« + *)<*> where c is the positive root of the equation x2 v2 1 y = 1 (11-33) b2 + c a2 c and a and b are the semiaxes of the contact ellipse; W is the load corre¬ sponding to the volume of the semiellipsoid of pressure, that is, W = %tt abpo, and v = (1 — vp2)/tE. By substituting this value of W in Eq. (11-32), we obtain $'(x) = ^2? _J_ „j) N i--* v' V d* N J0\ b2 + 4, a2 + <t>) V\/(l + <t>/a2)(b2 + Proceeding to the case of linear contact, we must insert o = » in Eqs. (11-33) and (11-32) and take into account that for this case W J- = liltb-pa Ry these means we obtain >! \ W I J- \ ( ** \ d* ~ + "■> i. - i- —*) viPTm and by setting x = x/b, <£ = 4>/b2, we obtain s (x) = 7- («1 + V2) / ( , I 7 - 1 ) ,-^-==== Ij Jo V1 + <t> /V<*>(i + <*>) = ^ (ki + Vi) lxy/x2 - 1 - ln (i + Vx2 - 1)] (11-34) The following expression applies for the half width of the area of contact 6 = 2 VWfa + Vi)p/L (11-35) where p is the effective radius.
Hydrodynamics of Rolling Elements 345 In what follows it will be more convenient to measure the thickness of the lubricant layer in units of Lh V = W(v i + l/2) (ll-36o) Thus we must change variables from h and x to 17 and x. It follows that in the presence of a lubricant we have v' = Vo = Ls' W(Vl + w2) Lho s = s' + ho v = v' + Vo = 2[i Vi2 — 1 — \n {x + y/x2 — 1)] (11-366) W(v, + Vt) Equation (11-31) now takes the form: (ll-36c) t+v’>V7W‘ or, with the new variables, dpo 0 /Tr - jT x 17' L2 dx 6,ll(t/1 +bi) (,' + *)» ir2(u, + v,)» Upon integrating this last expression within the limits from x = — 00 to x = —1, we find 6mi(Ui + U 2) dx (11-37) The integral in this expression is a function of 170. This function is denoted by 2(170) and may be determined by numerical integration; the relationship between 2(170) and 170 is given in Table 11-2. Table 11-2 0.1 0.2 0.5 1.0 2.0 5.0 Accurate values 2(170) 2.33 0.902 0.258 0.0986 0.0372 0.00982 0.0986 ijo-1-*71 2.34 0.901 0.254 0.0986 0.0380 0.0108 It develops that over the interval of change of 170 which is of practical importance the function 2(170) can be represented very accurately by the equation 2(170) = 0.0986170-1*376 (11-38) as may be seen from the above table. It follows from the above that 1
346 Theory of Hydrodynamic Lubrication By substituting in Eq. (11-37) the relationships (11-35), (ll-36c), (11-38), and (11-39), we finally obtain h0 = 1.13 [n(Ui + U5 ln a]0.727p0.364 [(TT/L)(w1 + U2)]0-0»1 (11-40) The mean thickness of the lubricant film at contact is determined from this expression by means of the external parameters, namely, the load per unit width of contact line; the elastic constants i>i and v2f of the rubbing bodies their effective radius p, and their speeds of movement Ui and U2; and, finally, the values m and a. It follows from the foregoing that for small values of Ui — U2 these last two parameters are the ordinary viscosity of the lubricant at atmos¬ pheric pressure and the piezo-coeffi- cient of viscosity, both at the tem¬ perature of the rubbing bodies. Note that ho is very slightly de¬ pendent on W and vi -f- v2. Equa¬ tion (11-40) is fundamental for calculations in practical cases with elastic rubbing surfaces. The lubricant film develops both pressures and shear stresses. On the surfaces of the disks without deformation, the pressures are so directed that they pass through the center of the disks. Consequently, their resultant forces W also pass through the centers. On the x axis, they go through the point Fig. 11-11. Frictional forces in disks. * = Jl'.*?***/ Jl'. vdx In this case, £ is negative. The two forces W have, therefore, no common line of action, refer to Fig. 11-1 la. On the surfaces of the disks, in addition to the pressures, there are shear stresses which we combine into the tangential forces Fi and F2, which are not necessarily equal. We assume Fi = F, + Fd and F2 = F, — Fd In these equations F, depends on Ui + U2 and Fd depends on U\ — U2. If now the forces W acting on the disks are combined with the con¬ siderably lower forces F* (Fig. 11-1 la), their directions change. The new resultants W have now a common line of action and with opposite direc-
Hydrodynamics of Rolling Elements 347 tions. From this follows the relation W | 2 p If we divide the height of the gap h in half, the points of bisection lie on a middle arc of the radius Rd, with 1 /Rd = %(\/Ri — l/R2). The common line of action passes through the center of this arc. Calcu¬ lated with respect to the centers of the disks, the torques are — fW — fW M[ = F.Ri = -sF- Ri and ATJ = FtR2 = R* Zp Zp We can also use £ to determine M[ and M'2. In addition to the determined resultants with a common line of action, we have yet to add the forces +Fd and — Fd. This changes the direction of the line of action, as indicated in Fig. 11-116. The torques resulting from +Fd and —Fd become M[f = FdRx and M" = -FdR2 Therefore the torques for both rollers become A/, = (F. + F„)ft, = ft, + Fdft, A/> = (F. - Fd)Ri = ft* - ftrfft* Zp We now define as friction coefficients / _ I ^ { = ^2 EjL (11 _ai Jl WRi 2p^W j2 WR2 2 p W K } In the cylindrical mid-plane there occur only the pressures and the shear stresses that depend on U\ — U2. The pressures furnish the resultant W through the centers of the middle arc; the shear stresses furnish the forces ±Fd. In order to determine fi and /2, we therefore start from this plane. For the deformed disks the pressures no longer pass through the disk centers. We divide the film thickness in the ratio of the displacements Si:s2. Since we have assumed the same modulus of elasticity for both disks, we have Si = s2 and we have to divide the widths of the gap in half. The middle arc placed through these bisecting points remains unchanged during the deformation. We, therefore, obtain in the respec¬ tive cylindrical planes the same conditions as on the disks without deformation: first the forces W through the center of this middle arc and then the tangential force +Fd for disk 1 and — Fd for disk 2. We can then determine from the torque at the disk centers the friction coefficients according to Eq. (11-41).
348 Theory of Hydrodynamic Lubrication We can now start with the calculation. From the developed diagrams p{x), we find £ = xp dx j J*'" pdx = xe xp d(x) j p dx = xe ^ xp dx and with x0 from Eq. (11-28): & [t = a vs /- - **di There are, furthermore, the shear stresses rd = p[(Ui — Ui)/h) depend¬ ing on U\ — U2' With the equation (aW \ . (otW \ 1 M = MO*-' = M. exp ^ P) and exp ^ pj = t according to Eq. (11-22), it becomes fx* EL fl 1 ^ = j- X*L dx = 1 - (n/Z/^D + s + A(x2 - I) In this way, we obtain for the friction coefficients * y]z^uUu\) - ~A /_.ip di U. - U2 1 _ /•• I U1 + (;2 (') v7B 1 — nf/fm„ho + s + .4 (i2 — 1) and /2 S 6Mu!+ u2) = ~AVF /_. Xp di dx Ui - Ui 1 f1 1 r __ Ui+ UiQ y/B J1 - nif/f^h0 + s + 4(i2 - 1) dx (11-42) The right-hand side terms of the above two equations are plotted in Fig. 11-12 for n = 0 and n = 0.8 as a function of zo/x, and on the left- hand side they are plotted as a function of n for xq/x€ = — 1. For a good approximation, a simplified relation that may be used to determine the coefficient of friction is given by f vUUt - (/*) ^mlnPO Note that the minimum film thickness and maximum pressure must be known and that the coefficient of fluid friction increases rapidly as Ui — Ui increases and may eventually exceed the coefficient of dry friction. According to some theories, dry friction takes place in the
Hydrodynamics of Rolling Elements 349 presence of a mono- or bimolecular layer of fluid. It would appear, there¬ fore, that when the coefficient of fluid friction exceeds that for dry fric¬ tion, there must be a breakdown of the lubricant layer. This break might be of a brittle or even plastic nature, but it does not follow New¬ tonian laws of fluid flow; furthermore, it would be limited to the region of particularly high pressures and would not affect our fundamental solution. We shall consider that in the presence of such breakdowns, the coefficient of friction would be equal or about equal to the coefficient of dry friction. In this case, we obtain good qualitative agreement between calculated values of friction and practical results. l 1.5 /? = 0.8 1 r —7—z~z—7—dx | 0.5- 1 -it 1 i i i n = 0.8 * 1 Q8 0.4 0 for-jr* = -1 *e -i-d.8-a6-a4-a2 6*£ Fig. 11-12. Graphical representation of the right-hand terms of Eq. (11-42). It has been shown that both viscosity variation with pressure and deformation of rolling surfaces have important effects on lubrication. An example will show these effects more clearly. The particular case considered in Ref. 6 is that of contact between two rollers that have curvatures of 1.475 in.-1 and roll together with a mean velocity of 206 ips under an applied load of 3,000 lb per inch of width. The rollers have a Young’s modulus of 30 X 10* psi and a Poisson’s ratio of 0.3. The lubricant has a viscosity m = noeap, where mo = 0.65 X 10“* lb sec/in.2 and a = 7.143 X 10“5 in.2/lb. Table 11-3 Surfaces rigid Surfaces deformable Lubricant isoviscous 0* = mo) Zimin = 3.74 X 10“8 in. W = 3,000 lb/in. Amm = 1.84 X 10-fl in. W = 3,000 lb/in. Lubricant viscosity varies (m = MoCap) hmin = 2.62 X 10"6 in. W = 477 lb/in. Amin = 4 X 10-6 in. W = 3,000 lb/in. Lubricant isoviscous (/x = /xo) hm\n = 4 X 10-6 in. W = 28.0 lb/in. Amin = 4 X 10"8 in. W - 47.8 lb/in. Lubricant viscosity varies (/x = /xoeap) hmin = 4 X 10“6 in. W = 40 lb/in. A0 = 4 X 10-# in. W = 3,000 lb/in.
350 Theory of Hydrodynamic Lubrication It appears, then, that for heavily loaded bearings it is almost essential to permit both the shape and the viscosity to vary. SOURCES 1. Martin, H. M.: The Lubrication of Gear Teeth, Engineering, vol. 102, p. 119, 1916. 2. Cameron, A.: Hydrodynamic Lubrication of Rotating Disks in Pure Sliding: A New Type of Oil Film Formation, J. Inst. Petrol., vol. 37, p. 471, 1951. 3. Weber, C., and Saalfeld, K.: Z. Angew. Math. u. Mech., vol. 34, no. 1/2, pp. 54-64, 1954. 4. Weber, C.: Boundary Deformation of a Half Plane by a Perpendicular Load, Z. angew. Math. u. Mech., vol. 30, no. 8/9, pp. 240-242, 1950. 5. Dorr, J.: Schmiermitteldruck und Randverformung des Rollenlagers, Ingr.- Arch., vol. 22, pp. 171-193, 1954. 6. Poritsky, H.: First ASLE Natl. Symposium on Fundamentals of Friction and Lubrication in Engineering, September, 1952. 7. Grubin, A. N.: Investigation of Contact of Machine Components, Tsentral. Nauk. Issledovatel. Inst. Tekhnol. i Mashinostroen., vol. 30, 1949. 8. Shchedrov, U.S.: On the Molecular Theory of Friction, Zhur. Tekh. Fiz., vol. 17, no. 5, 1947. 9. Davies, R.: Hydrodynamic Lubrication of a Cam and a Cam Follower Lubri¬ cation. Engineering, vol. 11, p. 37, 1955. 10. Sternlicht, B., P. Lewis, and P. Flynn: Theory of Lubrication of Rolling Con¬ tacts. ASME Paper 60-WA-286.
CHAPTER 12 INERTIA AND TURBULENCE EFFECTS INTRODUCTION Two of the assumptions underlying the derivation of the Reynolds equation are that the inertia forces of the lubricant are negligible and that the flow is laminar, assumptions 5 and 3 of Chap. 1. The familiar Reynolds number of fluid dynamics, given by pUD/p, expresses the ratio of the inertia forces to the viscous forces; in bearings a similar Reynolds number can be formulated by writing Re = pUh/p} where h is some representative film thickness. When this Reynolds number becomes sufficiently high, the two assumptions referred to above no longer hold. The inertia forces become of the same order of magnitude as the shearing forces, and laminar flow may give way to a turbulent or semiturbulent state. Under these conditions, the Reynolds equation as formulated in Chap. 1 no longer represents the true state of the lubricant, and we must return to the basic Eqs. (1-1) for a new approach to the problem. Inertia effects can be accounted for by including those terms of the Navier-Stokes equations which were originally dropped because of our assumption of negligible inertia. While the inclusion of these terms, the total derivatives Du/dt, Dv/dt, and Dw/dt, may complicate consider¬ ably the differential equations, there is at least no fundamental obstacle in the treatment of the subject. With the problem of turbulence, how¬ ever, there is a basic question of approach which merely reflects the general state of this branch of science. No definitive theory regarding turbulent flow as yet exists, and thus our solutions too will be based on differing hypotheses. It will be seen in the following pages that the contribution of inertia to the hydrodynamic forces in bearings is small. At very high speeds inertia effects may begin to have some influence, but then the motion of the lubricant becomes turbulent and our laminar equations, even though including the effect of inertia, no longer hold. The problem then must be solved on the basis of turbulent flow. Thus one can perhaps visualize three broad ranges of bearing operation: a lower region in which only viscous forces are of importance, a narrow intermediate range in which laminar flow prevails with both viscous and inertia forces present, and an upper region in which turbulent conditions prevail. 351
352 Theory of Hydrodynamic Lubrication EFFECTS OF FLUID INERTIA Significance of Inertia Terms Before proceeding with finding explicit solutions, let us first examine the hydrodynamic equations and determine at what stage the inertia forces merit attention. By rewriting Eqs. (1-1) for incompressible fluids and by making use of the continuity relation, we have (12-la) (12-16) (12-lc) By nondimensionalizing the above equations by writing du . u du , du , uv vd7 + 7ee+wTz+T dv , u dv , dv u2 dr r dd dz r dw , u dw , dw vTr+7ae+wTz -Ifp + afvtv + lto-A prdd p\ r2 dd r2) p dr p V r2 r2dB) -If + Uv'-w p dz p where dr2 r dr r2 de2 + dz2 r _ u w r ~ R U ~ W ~ hu _ z _ _ v p 2 ~h v ~ Th> p ~ KTfop and considering only the (r,0) plane, we have for the first two equations of (12-1) _ du . U du t _ du . UV _ _ 1 dp M r d^U . \ du . 1 d2U V df f d$ W dz f f dd pR2co [_df2 f df f2 dd2 . /R\2 d2u 2 dv ul /10 0 v + {h) eF+T>Te~T>J (12'2a) _ u2 _ __ dp p d^v JL dv , _1 d*v f df pR2u) [df2 f df f2 dd2 is <i2-2w _ dv , u dv , _ dv H— — -}- w — dr r dd dz Since (r,d) is the plane of bearing surface, z is in the direction of film thickness. By nondimensionalizing z and w in terms of h instead of R, we have made all velocity terms, as well as their first and second deriva¬ tives, of the same order. The relative magnitudes of the terms in Eq. (12-2) are thus given by their coefficients, the variables all being of order one. Since (R/h)2 is a very large number compared to one, the order of magnitude being 1,000,000:1, we can write for Eqs. (12-2)
Inertia and Turbulence Effects 353 _ du . u du . _ du . uv 1 dp . p d2u /in 0 N vaf+fM+wai + 7 = + (12'3a) _dv . u dv . . dv u2 dp p d2v . M ,,a? + FF9 + ",al-7= "a?+ (12‘36) The left-hand sides of Eqs. (12-3) contain all inertia terms; the last terms on the right-hand sides represent the viscous forces. Thus the inertia terms will be of the same order of magnitude as the viscous forces when their coefficients (the terms themselves being of the same order) are equal to the coefficients of d2u/dz2 and d2v/dz2J or pwh2 This can be rewritten as eM m eEh. m Re = * t (i2-4) MM ri which gives the magnitudes of speed, bearing size, and viscosity necessary to produce noticeable inertia effects in hydrodynamic bearings. This, however, does not mean that the contribution of inertia to the load capacity, or to any other bearing characteristic, is of the same order as the shearing forces. In fact, it will be seen later that, even when bearings operate in the range given by Eq. (12-4), the contribution of the inertia forces to the dynamics of bearings is still very small. For one-dimensional bearings the equation governing the flow of lubri¬ cant with inertia effects included is from Eqs. (1-1) given by Du _ dp . d2u p dt dx p dy2 (du du du\ dp d2u /10 >(u+uai + 9»i)~-te + 'lW' (12‘5) If only steady-state conditions are considered, du/dt = 0, and we have / du du\ dp d2u , n f,\urx + vTy)= -ix + liw* (12'6) Equation (12-6) contains all the assumptions leading up to Eqs. (1-3) minus that of negligible inertia. The expression — dp/dx + p d2u/dy2f instead of being equated to zero as in Eqs. (1-3), is set equal to the inertia forces of the lubricant. The exact solution of Eq. (12-6) is not simple. Two methods of approach will be formulated below, one an iteration process and the other a method based on averaged inertia across the fluid film. Both are justified on the basis that the contribution of inertia to the total hydro- dynamic forces is relatively small.
354 Theory of Hydrodynamic Lubrication Iteration Method Basically, this method consists of the following: the fluid velocities are first calculated with the inertia forces neglected, i.e., as a first approx¬ imation taking u and v as given by Eqs. (1-5) and the continuity equation and using them in the inertia terms of Eq. (12-6). Thus the left-hand side of Eq. (12-6) becomes a known function of x and y and the remaining unknowns are p and u on the right-hand side. Thus denoting by the subscript v the inertialess solutions, we have ( duv duv\ dp d2u /io7\ p{U’te+V’^-y)=-TX + ltW (12'7) where u = uv + uc p = pv + pc (12-8) uc and pc above are the correction terms due to inertia, and u and p without subscripts are the solutions for a bearing in which both viscous and inertia forces are present. If a further refinement is desired, the values of u and v can later be substituted into the left-hand side of Eq. (12-6) and a new set of u and p values obtained. By putting Eqs. (12-8) into Eq. (12-7) we have, since —dpv/dx -f- yd2uv/dy2 = 0, the equation P ( duv duv\ dpc d2ue n + v’-*j) = ~ di + liw (12-9) M = F{x,y) which in functional form is simply d2uc dy5 and can be solved for known sets of boundary conditions. The usual boundary conditions are u = U at y = 0 p = pa at x = 0 u = 0 at y = h p = pa at x = B where B is the bearing span. These conditions yield for the correction variables uc = 0 at y = 0 and y = h n2-10>) pc = 0 at x = 0 and x = B It also follows, since j* u(x,ij) dy = const, that JQh uc{x,y) dy = qc = const By writing y = y/h and referring to Chap. 3 for the expressions for an inertialess bearing, we have for the inertia terms «. § + *. ^ ™ [K,y + K,r + K,r + K<p]
Inertia and Turbulence Effects 355 where h' is the derivative with respect to x and Ki=- 6^ + 4 Kz = 18 - 36^° + 16 K2 = ~9 (x)2 + 27 7? " 14 = “9 (l)2 + 15 J ~ 6 In the above, the value of vv was obtained from the continuity equation du/dx dv/dy = 0. By integrating for uc from Eq. (12-9), we have Mx,y) = ^ ^ htyiy - 1) + y ^ (y* - 1) + ^ (y3 - 1) + §(r-D+f|(y6-l)] (12-U) By using this expression in f* uc(x,y) dy = <jn we have dPc 1 3mC72 T Kih1 Kzh' Kzhf K<h'] -di ~ - 12m'¥--r[-W+m+T5h + 2ih\ (12-12) By using the boundary conditions pc = 0 at x = 0 and x = B, the con¬ stant qe and the constant resulting from integration of Eq. (12-12) can be evaluated. Equation (12-12) will now be applied to both a plane slider and a journal bearing. 12-1. Slider with Inertia Considered. For a plane slider with the coordinate axis at the leading edge, we have h = hi — ax and thus by using one of the boundary conditions of Eqs. (12-10), we obtain dVc _ pU2(hi - h2) (Ko Ki K2 Kz KA dx B \h* ~h 2 h* (10/3)h ^ bh lh) K , „ 12 , , 2 . hi where *0 = _ Ao« _ _____ in_ It is seen that dpc/dx pU2/B. The sign of the pressure gradient is generally negative, and the effect of the gradient, therefore, is thus to reduce the velocity of the fluid; this is to be expected, because inertia acts to resist the acceleration of lubricant particles. Figure 12-1 shows a sam¬ ple velocity profile modified by the presence of inertia forces. By writing x = x/B, k = A./*, we Fig ^ Effect of inertia on velocity can use the remaining boundary con- profiles.
356 Theory of Hydrodynamic Lubrication dition to integrate Eq. (12-13) and thus *<■> -[{[ - (r^)’1 (‘ - T=~m)+ fln r=V4 <12-14) 3k — 35 2k It is seen that pc PJJ2 and that it is independent of the inclination of the slider or film height but depends only on the ratio hi/8. The value of this pc is always positive, thus adding to the total load capacity; its shape is similar to pv. By writing for a modified Reynolds number Re* = ^ (12-15) we have for the total pressure p(x) = pv + pe = (pv + Re* pc) (12-16) 6^2 x(l ^c) where pv = 2k — 1 (k — x)2 *S so^u^on ^or a s^^er without inertia and pc is the expression given in the brackets of Eq. (12-14). By inte¬ grating for load capacity, we have W. = LB // pc(x) dx = He* We = pU'BLWc hi , t77 1 6 k2 — k 1 k2 — k, k /10 where W< ~ 7 35 (2k - l)2 5 2k^~l F^l ( ) The total load capacity is w = W, + We = (W. + Re* We) (12-18) where Wv = 0/c2 (hi Values for Wv and Wc are given in Table 12-1. At Re* values of 5, the correction is of the order of 10 per cent. The curve We has a maximum at k — 1.65 which compares with a maximum at k = 1.85 for Wv. The location of the resultant load through the slider is essentially unaffected by the inertia term. For friction dn.
Inertia and Turbulence Effects 357 Table 12-1. Inertia Effects in Thrust Bearings k = hi hi — hi 1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2.0 2.5 3.0 Wv 00 5.330 3.130 2.190 1.680 1.330 1.110 0.940 0.818 0.719 0.634 0.405 0.302 0 We 0.1429 0.0857 0.0588 0.0449 0.0349 0.0283 0.0237 0.0198 0.0167 0.0142 0.0123 0.0069 0.0044 0 F, 5.050 3.460 2.760 2.348 2.095 1.920 1.789 1.685 1.604 1.543 1.360 1.272 1 Fe -0.0286 0.0097 0.0287 0.0384 0.0431 0.0452 0.0462 0.0458 0.0451 0.0441 0.0428 0.0368 0.0316 0 0.820 0.948 1.105 1.260 1.400 1.572 1.730 1.901 2.080 2.260 2.440 3.360 4.290 •0.0148 ■0.0105 0.0066 0.0029 0.0003 0.0026 0.0045 0.0063 0.0074 0.0082 0.0099 0.0095 and upon integration r,(£) dx = Re* Fc = PU*htLFc h i or F = (P, + Re* Fc) ki Fc = ^ 18 fc - 1 where ^ 1 35 k 35 (2k - l)2 and the coefficient of friction is (12-19) (12-20) (12-21) / = ^ =/,(! + Re* /.) where / = *ji - Je f9 wv (12-22) and _ hi J_ (2 ln [k/(k - 1)] - 3/(2k - 1) Jv B 3k ln [k/(k - 1)] - 2/(2k - 1) Values of FV) PCf fvy and fe are given in Table 12-1. In general the friction is increased except at very low values of k. 12-2. Journal Bearing with Inertia Considered. For the inertialess component of the final solutions, we shall here employ the so-called half- Sommerfeld expressions, i.e., bearing characteristics based on integrations over the range 0-tt, with the negative pressure of the region t-2tt set equal to zero. These expressions are from previous considerations (Chaps. 3 and 7).
358 Theory of Hydrodynamic Lubrication nUR _ , _ 6c (2 + € cos 6) sin 0 Pv = -^Pv where p. = ^ ~ (f+t cos 9)* 1 12t*2 1 . 67T2C tt COS 4>v = 71 ,wo , 57 TT Sin <f>v = S, " (1 - C2)(2 + €2) *Sr (1 - 62)^(2 + €2) T /l - €2\^ tan <f>v = |f—LJ (12-23) „ _ * , fr 2t 1 + 2«2 * C " re q _ C2)M» 2 + «2 / = where f = (1 + ~ Jv RSJV wnere^2(1 _ ^ + ^ The expression for the correction pressure gradient is similar to that of Eq. (12-13): dVc \A h'(Kx , ,A'3_lAA] ~dd = pl [F3 ” ¥ \ 2~ + To73 + T + tJJ (12“24) with 4 [36 € 2 € 1 1 , 1+el " [ 5 (2 + e2)3 5 (2 + €2)2 3 2 + c2 1 - ej and h, of course, given by h = C(1 e cos 0). The pressure correction after integration is Pc = ~ Re** pc = PU*p. (12-25) where . 216 e (h /2\ 12 € //, I3\ Ve 35 (2 + e2)2 y7r 2 ) 35 2 + <2 \t 2 ) _2lnl/i+lln 1+. 7 X t 7 1 + < cos — «2)w sin 9 2 + t2 1 + f cos 9 « (1 — «2),c' sin 9 2 2 + «2 (1 + « cos ey- r ! Li e i 1 2 [1 + X tun2 (9/2)]2 1 -I- X tan2 (9/2) ^ j = , 1_ x = Ijtj 3 1 + X tan2 (9/2) 1 + t and Re** has been redefined as 25 (I)’ (12-20)
Inertia and Turbulence Effects 359 Thus the total pressure in a journal bearing with both viscous and inertia terms considered is p = (?• + Re** ?•) (12-27> In general, the correction pe is positive except at low eccentricities and high values of 0, when it becomes negative. For the two load com¬ ponents, we have cos <f> = J- cos <f>v + Re** (12-28) where after integration of Eq. (12-27) 1 432 c(l + €2)* /0_ . 24\ 1 €(1 - €2)^ -(27' + v)i1 irSd 35ir (2 -j- «2)3 \ tt / 35 (2 + €2)2 3tT (1 — €2)* - (1 - €2) . IT 1 — (1 — €2)* + 35 €(2 + c2) + 7 e 4 (1 c2)l* ^ 1 4~ € (12-29) 7ir (2 + €2) 1 1 - € (12 and 4 sin 0 = 4" s*n <£*> + Re** — /S Sv 7tSc2 where 1 324 c2 18 e2 (j 1 2 7rSc2 35 (2 + e2)3 + 5 (2 + e2)2 35 2 + e2 + 7 _ 1 1 - 62 1+J 5 6(2 + 62) 1 - € The total load capacity by neglecting terms of higher order is with 1 _ (1/SV) sin <f>v (\/irSc2) + (1 /7T<Sp) cos <f>v (1/^ci) irSc 1JSl The new attitude angle is <t> = <f>v + Re** 0C (12-31) „.u j. (!/&„) COS (f>v (1/mSU) - (l/iSr) sin 4>v (1/tt*Sci) where 04 (Ijsj> Values for all the dimensionless constants are given in Table 12-2. It can be seen that the effect of the inertia forces on the total load capacity and attitude angle of a journal bearing is quite small.
360 Theory of Hydrodynamic Lubrication The correction factor for friction is Fc = Re**Fc = pU'-LCFc (12-32) , * I-108 1 - e2 18 1 - e2 2 1 6 e * [ 35 (2 + f2)3 35 2 + e2 35j and the friction coefficient / = /»(! + Re**/c) (12-33) with /. = '• - m. F. 1/S, This correction coefficient fe is quite small and for t > 0.5 becomes nega¬ tive The correction friction force Fe becomes negative at € > 0.7. Table 12-2. Inertia Effects in Journal Bearings e 1 “5" cos <t>v Top 1/t Set ;ksin +' 1/x5c2 l/*Sv 1 /*Sc <t>*> deg <t>c, deg 0 0 0 0 0 0 0 90 0 0.1 0.0603 -0.0075 0.9425 0.0003 0.945 -0.0002 86.34 0.46 0.2 0.2451 -0.0107 1.8861 0.0019 1.902 0.0005 82.67 0.33 0.3 0.5679 -0.0138 2.8361 0.0069 2.893 0.0041 78.68 0.26 0.4 1.0582 -0.0125 3.8087 0.0183 3.954 0.0169 74.47 0.22 0.5 1.7778 -0.0055 4.8369 0.0377 5.154 0.0334 69.82 0.19 0.6 2.8602 0.0104 5.9903 0.0669 6.638 0.0648 64.48 0.17 0.7 4.6303 0.0387 7.4206 0.1061 8.748 0.1105 58.04 0.16 0.8 8.0808 0.0847 9.5200 0.1545 12.490 0.1726 49.67 0.16 0.85 11.4760 0.1216 11.2741 0.1820 16.090 0.2142 44.49 0.15 0.9 18.2056 0.1626 13.8500 0.2118 22.880 0.2576 37.26 0.15 0.95 38.2694 0.2288 19.7583 0.2452 43.070 0.3158 27.31 0.15 1.0 00 0.4488 oo 0.2857 oo 0.3700 0 0.15 Method of Averaged Inertia Since the inertia forces are small and the fluid film is very> thin, a reasonable approach is to average out the inertia effects across the film. We can thus rewrite Eq. (12-6) in the following manner: fl fh I du du\ , \ dp d2u noo4\ ~^+M^ ( 3 ) The left-hand side of Eq. (12-34) after integration is a function of x alone, and the expression can thus be integrated for y in the manner of previous solutions. By writing p fh ( du . du\ 35). (“S + Vj
Inertia and Turbvlence Effects 361 we have from Eq. (12-34) 0 = /(*> (12-35) It will be recalled that this approach was used in Chap. 5 to obtain an expression for the pressure in gas bearings, and Eq. (5-28) is based on Eq. (12-34). No explicit solutions of Eq. (5-28) were possible because of the variable density which made the equations nonlinear. However, here, for incompressible fluids, it is possible to directly integrate Eq. (12-34) and obtain solutions for both thrust and journal bearings. By integrating Eq. (12-35), we obtain u = }4f(x)y2 + Ci(x)y + C2(x) By using the continuity equation du/dx + dv/dy = 0, we have v = - f&dy or v = - + C[(x) £ + C'2(x)j/] + C,(z) where the primes denote differentiation with respect to x. The boundary conditions in general are u = 0 v = 0 at y = 0 u = U v = V at y = h When the first three conditions are used in the expressions for u and t>, we have u = \f(.x){y'-hy)+^ (12-36) > = - (J/W + i [x - l/Wh] I) (12-37) We can now use the remaining boundary condition v = V at y = h in Eq. (12-37) to obtain the value of f(x) and finally of dp/dx. These expressions are /(x) -»/,*+ «/«,(?) + £ - (12-38) g-«K») + p[-JlW'-£vWK + il/*f - + ±f(x)Uhk' + ^/’(z)^'] (12-39) Equation (12-39) can be evaluated when (7, V, and the function h are given. This expression will now be applied to thrust and journal bear-
362 Theory of Hydrodynamic Lubrication ings starting with the simple case of squeeze film action between two plates. 12-3. Squeeze Films. Considering two infinitely long parallel plates of span B approaching each other with a relative velocity F, we have (7 = 0 h! = 0 V = const and so from Eq. (12-39) 2-('+¥)*> Since /(x)=g f yrfx + g-^ + g we have p(x) = J* (m + + £f) dx The integration yields p(x) = (m + x) (12-40) Equation (12-40) differs from the standard solutions given in Chap. 7 by the supplementary term phV/5 in the parentheses, which does not depend on the viscosity but which contains the density. The total correction is seen to be proportional to the square of the approach velocity V2; its magnitude will not be large owing to the value of h, which is of an order much lower than all other terms in Eq. (12-40). 12-4. Journal Bearing. For ordinary bearing operation, expression (12-39) can be somewhat simplified. Under conditions of steady loading and constant linear velocity, we have F = 0 and U' = 0, and Eq. (12-39) becomes t - ««*>+<• [sL'(j)+ + TS™*'*'] Returning to expression (12-38), we have after integration for our conditions ,, . 6 Uh , C\ /(x) UT + ¥ which can be rewritten as /(x) = ~ (ht - h) (12-42) where hx is the film thickness which, according to Eq. (12-35), makes d2u/dy2 = 0 and is essentially an integration constant. When Eq. (12-42) is used in Eq. (12-41) we obtain dp 6 yV - = — {hx-h) + 30h [ \/i/ (12-43)
Inertia and Turbulence Effects 363 The first right-hand term of Eq. (12-43) is seen to be similar to Eq. (3-5); the remaining term is that due to inertia forces. If we write pC2 _ pUC2 __ Re /C\ P 6nUR 6pft 6 \R/ the expression for the pressure gradient becomes dp (hx - h)C2 dd hz ai(S)Ki)’-4 (12-44) By writing pv for the Re = 0 case, i.e., with the inertia terms neglected, we have for the solution of Eq. (12-44) p = pv + \pe hx = h0 + \he By substituting these last expressions into Eq. (12-44) and neglecting small quantities, we have dpv _ (h0 - h)C2 dd h3 dpc hcC2 The values of pv and h0 are the known Sommerfeld solutions: _ _ e sin 6(2 + e cos S) L _ 1 — c2 p' ~ (2 + e2)(i + e cos ey 0 ” 2 + «2 By using the same Sommerfeld substitutions and the proper boundary conditions on pe, we obtain he r 2 + e2 , e -f cos 6 Pc = ~nT\ ini —77— cos"1 C(1 - e2)>* I 2 1 + € cos 6 «(1 — e2)**(€2 — 3e cos 0 — 4) + 2(i +«cos ey sin in e j , 3/1 - «2Yr 1 1 1 , 2 1 + t + 5 \2 + «7 [(1 + «)2 (1 + t cos 0)2J + 15 1 + « cos e (12’45) u i, _ 4(?(i -<2)* r 6* 11 1 + ‘i where hc 5,(2 + e2) [(2 + «2)2 3 1 - t J The ratio of the pressures obtained from the purely viscous and the viscous-inertia solutions is given by -?■ = 1 + x — p. p. or, by using the proper relations p=l+ile^ (12.46) Pr 6 Rpv
364 Theory of Hydrodynamic Lubrication For an eccentricity ratio of 0.2 the value of pc/6pp, which is a function of c and 8 only, ranges from 0.009 at 8 = 30° to —0.01 at 0 = 150°. Now, from the criterion for turbulence given in the second part of this chapter, the maximum value of Re at which laminar flow still prevails is Re = 41.1 (R/C)M. Thus Eq. (12-46) can be written (p.)-.. 1 + 6p. 411 (ft) By using a representative value of 0.001 for C/R we get for a maximum possible correction 41.1 = 0.01 • 41.1/VpOO = 1.3 per cent which, as previously mentioned, is relatively small. 12-6. Slider Bearing. By using for the slider film thickness the expo¬ nential function h = hmtor-*IB where r = hmiu/hh and writing T = — ft — ^°lin X — in hmia p B P 6nUB A 6nB 6B K we get for the pressure gradient from Eq. (12-43) (12-47, By proceeding as above with a solution that consists of pv and a correc¬ tion pe due to inertia, we have dpv _ (h0 — ft) ft*,u jjf2 dx h8 dpc_hchlia lnrf ApV I dx /i3 30 [ \h) J The two equations integrated with p = 0 at x = 0 and x — \ yield 1/1 - r32 1 - r22\ r V1 - r3 1 - r2 J /I - r2\2 . , 2 ln r/_ 1 - r”\ (l - r3) ?,+ 15 V 1 - r3) . 27^/1-^’ (12‘48) Pc = 40 By writing for the load capacity
Inertia and Turbulence Effects 365 we have after integrating pv and pe (\ - r-^i i - b = r2 - 1 I 3 ln r _ 2 ln r I 2 In r \ 1 — r8 1— r2/ b _ 27 In r/1 - rJV . . ln r f\ o 1 - (r3 ~ l)/(3 ln r)] P‘ ~ "lo- \T=7>) p' + 16 L T^75 J and the ratio of the load capacities is (12-49) T, = l + K^B£= l+KjgRe (12-5°) where K is the ratio of Pc/Pv and is given by „ _ 27 ln r /l - r2V 40 V “ r7 _ „ 1 - (r3 - l)/(3 ln r) i ^ ^ 1 /10C1\ + 15(r2 - 1) 1 - (r3 - l)/(3 ln r) 1 - (r2 - l)/(2'Inr) (12'51) 1 - r3 1 - r2 which is seen to be a function of r only. Here again by numerical examples it can be shown that even at high linear speeds the inertia effects on load capacity are of the order of 1 per cent or less. ACCELERATION EFFECTS IN BEARINGS Tangential acceleration of journal or thrust runner always occurs in starting and stopping of engines, and some machine components such as piston rings or spur gears experience a cyclic sequence of acceleration and deceleration. The following is a simplified examination of the effects such accelerations have on the performance of bearings. We shall in Eq. (12-5) ignore the inertia of the lubricant and consider only the inertia term due to the unsteady linear velocity of the bearing surface. Thus the equation we shall consider is du dp d2u pTt = -d-x + ^W* (12'52) When this is integrated with the ordinary boundary conditions, we obtain d£-wk-^-lw which is an equation identical with (3-5) except for the acceleration term
366 Theory of Hydrodynamic Lubrication pdU 2 yields 6pUR € sin 0(2 + e cos 0) V - Va = 2 dt integrated over 2ir with p = pa at 0 = 0, Eq. (12-53) C2 (1+6 cos 0)2(2 + €2) 6 sin 0(e2 — 4 — 3c cos 0)(1 — e2)>* dt/ + ' 2(l+e cos 0)2(2 + c2) p df + COS’ / cos 6 + « \ pR dU pR dU a /10 (r+tco8f»j^- dT® (12'54) The last three terms of Eq. (12-54) are the contribution of the accelera¬ tion term to the bearing pressures. These tend to reduce the magnitude of p. The reduction increases with the value of the acceleration. As in the steady-state case, the term W cos <f> yields zero, indicating a locus at right angles to the load. For the load capacity then with sin <f> = 1 f2w 2r f2w fo W = L I p sin 0R d0 = — RLp cos 0 + RL / cos 0 d0 Jo o Jo d0 which when integrated results in 2-KhpR'LN v2tR*Lp dN 5C2(2 + c2)(l - 62)* 10(2 + c2) dt The percentage decrease in load capacity due to an acceleration dN/dt is from Eq. (12-55) AW _ 1 dN/dt C2(l - €2)* W p/p N 4 which shows that the reduction is proportional to the relative accelera¬ tion dN/dt/N. For slider bearings with a film given by the equation h = h2(a — ax + x) where x = we have £> dp = 6 pUB I" 1 Ci 1 pB dU n2KfiN dx hi2 [(a — ax + x)2 h2(a — ax + x)3J 2 dt By use of the boundary condition p(0) = p(l) = 0, the integrated pres¬ sure is 6pUB x(a - 1)(1 - x) P{X) h22 (a + l)(o - ax + x)2 + 2{J - 1) It [(o - ax + xY + ~ + ^ ] (12'57) The second right-hand term is always negative, thus a positive accelera¬ tion will yield lower pressures than for steady-state conditions.
Inertia and Turbulence Effects 367 The load capacity by integrating Eq. (12-57) is <«•> and the relative reduction in TT due to the acceleration is ATT 1 dU/dt hf(a ~ l)3 n2 . TT m/p U 24(a + 1) ln a - 48(a - 1) which here too is seen to depend on the relative acceleration dU/dt/U. The reduction of load capacity referred to above is to be under¬ stood only as an instantaneous effect. As the acceleration per¬ sists, the velocity will increase and the load capacity will rise with time. The above simple analysis is based on the assumption that the acceleration within the lubricant is proportional to the distance from the stationary surface. When no such assumption is made, the analy¬ sis given'in Ref. 7 shows zero change in load capacity (as opposed to a decrease) for an initial very brief period of time, after which, for a constant acceleration, the load ca¬ pacity rises nearly linearly with time. The relation between load capacity and time for the case of a = 2 is shown in Fig. 12-2. EFFECTS OF TURBULENCE When bearings are operated at sufficiently high speeds or high clear¬ ances or when the viscosity of the lubricant is sufficiently low, the flow will change from laminar to turbulent. One can visualize the inception of turbulence as that instance in which the centrifugal forces become high enough to overcome viscous resistance and thus, instead of the laminar streamlets, circulation and vortexes are set up. A turbulent fluid film affects the performance of bearings in a number of ways: it raises the power losses, lowers the flow, and alters the locus of shaft center. Before a study of turbulence in bearings can be undertaken, it is necessary first to understand the conditions that promote turbulence and to deter¬ mine the range in which either turbulent, intermediate, or laminar flow prevails. 4 vt/hf Fig. 12-2. Relation between load capacity and time for an accelerated slider.
368 Theory of Hydrodynamic Lubrication 12-6. Criteria of Fluid Instability. The theory of stability of fluid films between rotating cylinders is due to the classical work of Taylor. If U is the laminar tangential fluid velocity, then by referring to Fig. 12-3 we denote by U + uf v, and w the three velocity components of the disturbed fluid. The three turbulent velocity components are assumed to be given by the following expressions u — U\ cos \zeai v = Vi cos \zeai w = W\ sin Xze** (12-60) where wi, v\, and Wi are functions of r only. These velocities when used in the appropriate Navier-Stokes equations are then analyzed for the value of <r. If <r > 0, the disturbance velocities increase indefinitely with time and turbulence prevails. If <r < 0, the tur¬ bulent velocity components die out with time and the fluid returns to stable operation, a = 0 is then the threshold of fluid instability. The value of a will, of course, be related to the parameters of geometry and bearing operation, and thus physical conditions are established for the onset of turbulence. The value of a can in general be either real or complex. It can, however, be shown that a cannot be imaginary or complex unless u\, Vi, and W\ are themselves complex. Furthermore, if the complex roots for a have a real part, it would imply an oscillating disturbance of increasing magnitude. This has been shown by Rayleigh to be impossible for compressible fluids, and Taylor’s experiments did not detect such a tendency in liquids. The evaluation of <r will thus be restricted to real roots with the motion being stable or unstable according to whether <r is negative or positive. * Our interest lies primarily in small clearances, and thus by eliminating all terms C/R\oi order higher than one, we obtain for a stability criterion the following expression Fig. 12-3. Coordinate system for concentric cylinders. M = ir*p2(Ri 4" Rz) (12-61) 2a>12C8#i2[l - fl(fl2/fli)2](l - 0) where 12 is the ratio of the velocities of the two cylinders, or w2/a>i. Three possible cases arise. Case /, 12 « 1. This represents the case of cylinders rotating in the same direction with slightly differing velocities. For this case 1+12 M = 0.0571 1 - 12 or 7TAV2{R\ + &) 2a>i2fli2C3[l - 12(tf2/fli)2](l Eq. (12-62) tells us the following: (i+i) ‘ 0 0571
Inertia and Turbulence Effects 369 1. If w2/wi < (/?i/#2)2, the motion is unstable for any value of an higher than that obtained from Eq. (12-62). 2. If w2/a>i > (#i/#2)2, the motion is always stable (c*>i is imaginary). Lord Rayleigh’s criterion for the stability of inviscid fluids agrees with statement 2. For part 1, Rayleigh's criterion was that the fluid will be unstable at all speeds, while here instability will set in only above an given by Eq. (12-62). The type of instability formed is periodic along the length of the cylinder with a wavelength almost equal to the dia¬ metral clearance, or 2t/\ = 2C. Case I Ij 0 < 12 < 1. This case covers any velocities of the two cylinders, provided the two rotate in the same direction. Two groups of solutions arise: (1) C/Ri « 0. This is the case applicable to bear¬ ings where the clearance is negligible compared to the diameter. The solutions here are only slightly different from those given for case I, the correction being of the order of 1 per cent. The value of M is here given by M = 0.0571+ 0.00056 (12-63) 1 \l 1 "j" (2) If (C/Ri) is not negligible, the value of M is M = 0.0571 (i±? - 0.0652 §-) + 0.00056 (i±| - 0.652 £)"* (12-64) Case IIIj 12 < 0. When the cylinders rotated in opposite directions, the relevant equations could not be solved explicitly. Two specific cases were computed and gave the following results Ri, cm Ri> cm ft XC/t M 3.80 4.035 -1.50 1.73 0.00134 3.55 4.035 -1.347 1.73 0.00134 To sum up, the values of M given above with M defined by Eq. (12-61) set up quantitative criteria for the onset of instability. In general, when the inner cylinder is stationary, the flow is always stable; when the inner cylinder rotates, instability will set in at some definite speed. This is not yet full turbulent flow. At this point, orderly vortexes are formed, and while these persist, we have an intermediate state. Only when these vortexes break up into an apparently random pattern do we enter the fully turbulent region. These vortexes can be mapped by a study of the fluid streamlines. For the case when 12 « 1 and C/Ri « 0 the streamlines are given in Fig. 12-4. Their spacing gives an estimate of the velocity of the fluid
370 Theory of Hydrodynamic Lubrication at any point. The circulation in a section of the fluid consists of a series of nearly square vortexes rotating in opposite directions. For the case of 12 = —1.5 the flow is divided into two separate regions as shown in Fig. 12-5. The region adjacent to the inner cylinder is similar to the previous case. The outer region is much less vigorous and rotates in a direction opposite to the inner vortexes. I neiAr ^ulin^Ar Hnfar /*u tin War 'Inner cylinder Outer cylinder 'Colored 'fluid Fig. 12-4. Taylor vortexes between con¬ centric cylinders rotating in the same direction. Colored fluid Fig. 12-5. Taylor vortexes between counter-rotating concentric cylinders. If the effect of eccentricity is neglected, Eq. (12-63) above is immedi¬ ately applicable to journal bearings. By replacing M in Eq. (12-63) by its proper value, we have xV(fti + Rt) 2u)i2Ri2C3[l - U(Ri/R*)*\(l - 12) For normal bearing operation = 0.0571 + 0.00056 1 12 1 - 12 1 + 12 and so Q = ss = 0 0>1 -V2 R Ri = R\ = 0.0577 2«i 2R2C3 This yields as a condition of laminar operation the inequality 41.1* w < C*R» (12-65)
Inertia and Turbulence Effects 371 where w is the angular velocity of the journal. In terms of the bearing Reynolds number this can be written as Thus to keep the bearing in the laminar region, the viscosity must be high and clearance and diameter low. As mentioned at the beginning of this chapter, there is not as yet available a formal treatment of turbulent flow. In the following pages, therefore, each case proceeds from a different hypothesis, dictated either by mathematical necessity or the investigator’s preference. However, these analyses do possess one common feature, and that is the postulate that the pressure profile is essentially proportional to the square of the mean velocity. 12-7. Turbulent Operation of Journal Bearings. If on the basis of short-bearing theory we neglect the pressure variation in the circumferen¬ tial direction and assume the flow to be a simple Couette flow determined by the relative motion of the surfaces only, we can write The flow in the axial direction is taken from Blasius’ law of friction, which states where/i = 0.326, n = 0.25, and w is the average velocity in the z direc¬ tion. Thus (12-66) Qz = uhdz = -g- dz (12-67) From the continuity equation ^ («.) dx + (0.) dz = 0 or d_ dz where K fivnP 22+" Integration of the preceding expression yields /n \ 1 (U C \2_n (sin 6)2~n [ri . n . p(0,z) - (3 _ n)K \2 Re) h3 ^ 2 ^ + ^ ^
372 Theory of Hydrodynamic Lubrication By use of the boundary conditions ^ = 0 at z = 0 p = 0 at z = ^ v(ez) = I /j/V-w_L J*sin eY~n. lYA3- - *3-l (3 - n)K\2R) Cl+n (1 + « cos d)8 [\2) J (12-68) In integrating expression (12-68) for the two load components use is made of the following formula: •/; <«*»+• c°s *)-* de=^ ?(T+W)F ^ *+b (j where T(x) is the gamma function and F(a}b; c; a:) is the hypergeometric function. For the hypergeometric function we can also write F [I Mr: *+ (rr^y] = (1 + z'yF{t’1 ~s + s + z2) The integrated load components then are w A 1 wc*v-n(L\~n 1 PVl I)l~n K 4-n \2j C3 B \ (UCe)2-" (L\~» \ y I)l~n K 4 - n \2J C (12-69) with C3 3 — n A /- 3« r 2 /5 0 6 - n 2\ ^ -r=7i * (2’2; —5e) 2 4-n 2 The resultant load capacity is P = (UCt)2 /L\4 " r LI)'-»K(4 - n)C*\2j K ^ ; With Re = CU/v, (R\ (kYn Re,-. = >«(* ~ ") I (12-70) P \c) \d) Kt /lire2-'1 (A2 + B*)* (lZ iU> The attitude angle is of course given by <f> = arctan-^ (12-71)
Inertia and Turbulence Effects 373 It will be seen that here in addition to the Sommerfeld number and the L/D ratio, the Reynolds number is also a parameter; i.e., the attitude of the shaft is determined by the dimensionless grouping 3-n Re*-" = /(«) For laminar flow n — 1, and thus this grouping will be reduced to the standard form obtained in Chap. 3. For laminar flow K = 96 and then 2 4 6 8 10 12 14 16 18 10 20 30 40 50 60 70 80 90 4 0 80 1 20160 200 240280 320 5U/0)2 75tf*°75 1.0 0.8 0.6 f 0.4 0.2 0. A* = 2,000 L-4,000 *"6,000 10,000 20.000 M ion i/nor _ i r 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 SU/O)2 Fig. 12-6. Load capacity of journal bear¬ ings under turbulent conditions. Fig. 12-7. Locus of journal center under turbulent conditions: A, turbulent; B, laminar; semi¬ circle. A = 2/(1 - c2)2, B = (tt/2)(1 - e2)-* and Eqs. (12-68) to (12-71) will reduce to those of laminar flow. Figures 12-6 and 12-7 show the charac¬ teristics of turbulent flow as a function of the grouping 5(L/Z))2*76 Re0-76. As before, it is seen that turbulence in the fluid film causes an increase in load capacity. The position of shaft center, however, as seen from Fig. 12-7, is to the right of the laminar locus yielding higher values of <f> for a given eccentricity. 12-8. Turbulent Operation of Slider Bearings. For the study of slider bearings we shall employ the principles of Prandtl’s mixing length theory. This concept introduces a length similar to the mean free path in the kinetic theory of gases such that a certain quantity of the turbulent flow is assumed to be preserved throughout this mixing length. By represent-
374 Theory of Hydrodynamic Lubrication ing by a bar mean values and by subscripts the turbulent fluctuations, we have u = u + Ui w = w W\ v = v + vi p = p + pi The Navier-Stokes equations for incompressible fluids and one-dimen¬ sional bearings are from Eqs. (1-1) / du du\ dp + = ~di + ,lVu ( dv dv\ dp _2 p{UYx + Vry)=-Ty + >iV2v where V2 is the Laplacian operator. When the expressions for the turbulent velocities are introduced in the above equations and averaged, these equations become ci- d2U d2U , d*v dH s,nce aZ^W' and «W dUi2 , diiiVi „ dUiVi ~~z and —— « —— dx dx dy d2u d2v dUi2 dU\Vi _ paF = p^ = p^F = p-yr = ° (.du ... diZ\ dp d ( du _ _ \ (uTx + vij)--Tx + ai\!lai-'m'v') (.dv dv\ dp d ( dv A p{u^ + VTy)= -Ty + aTj^Vy-^) — _L — — 0 dx + dy “ By calling the mixing length I and assuming that the transferable fluid quantity q is a function of y only, we can write for the fluctuating value of q over the length I l«i| = q(v +1) - g(y) By expansion of q(y + I) in a Taylor series, we can write and we have (12-72)
Inertia and Turbulence Effects 375 which for very small values of I gives The momentum of flow will here be considered the transferable quan¬ tity, and we write M = (12-73) The fluctuating velocities u\ and Vi may be related to the mean velocities by use of the mixing length I. Since dv/dx <C du/dy, we can write = -pivv'=p^gy or more appropriately, since du/dy can change sign r = pi2 fy <12-74> du dy Two more assumptions are made. One is that the turbulent friction is much larger than laminar friction, and thus p d2u/dy2 = 0; the second is that, by analogy to pipe flow, we can write for the slider pi2 = k2y(h - y) (12-75) where k is an experimental constant. With the expressions of r and I used in the first of Eqs. (12-72) we have (. du du\ dp d [72 . du\du'\ /10 p{UK+VTy)= -fx+TV[ky{y-h) Ty\Ty\ (12'76) By making y an order 8 and du/dx an order 1, we have for Eq. (12-76) the following orders of magnitude \ dx dy) dx dy \_ \dy\dy\ 8 8 L 8 8 J The left-hand side is thus seen to be of an order much lower than the right-hand side and is here neglected. Thus we have which is the pressure equation for turbulent flow in bearings. Here dp/dx is written in place of dp/dx on the assumption that v and V\ are small and that consequently there is no pressure variation across the
376 Theory of Hydrodynamic Lubrication film. By integration of Eq. (12-77), - »> (fj)* = C'W - fx y If we omit the small region of back flow, then du/dx is always negative, and by writing dp/dx = p, we obtain (12'78) Since y(h — y) > 0, we must also have C\ — py > 0. Equation (12-78) can be used in its given form, for it can be shown that, if the region of back flow is omitted, the equation retains its form, i.e., has a negative sign for regions of both positive and negative pressure gradients. By making the substitution i =si"2 f we can write the integral of Eq. (12-78) as “ = - 2~nr IX1 ~ Ssin2 0 dt + Cl Since the term {ph/C/) sin2 f < 1, we can expand the integrand into (i _g1 - l.l.^sinu+ • • • and, taking only the first two terms of the expression, -fe--')*** The integration, with y/h substituted back for sin2 yields * - -2-# [(■ - w) J + k <12-™ By using the boundary conditions u = V at y = 0 u = 0 at y = h we obtain k*U* C'-u c,= 2t2 [(* + mr*p) + {1+wm*) ] The flow as given by q = j* u dy is « = Uh - {(l - $) foh sin- ^ dy + 4g-1 Vu(h-y)dy
Inertia and Turbulence Effects 377 Since sin-1 dy = | and y/y(h - y) dy = | h- '-"-tI'-II «,2-80) From continuity requirements dq/dx = 0, and so which upon integration yields 5h + C, = ^ (l + ~ pj! where C3 is an integration constant. By squaring and rearranging the above, expression, and remembering that p = dp/dx, we have dp kHJ2 dx T2 By remembering that h = h2[a — (a — 1 )x/B) and using the boundary conditions of p = 0 at x = 0 and x = B, we integrate Eq. (12-81) for the mean turbulent pressure and obtain k2U2B | , /, a — 1 _\ ■ x(\ — tx) \ MO QON P = t) \r In (1 - — - xj + . (-—- D^p) d2-82) Tr2h2a2{a where x = 4s r = 24a2 a(a - l)C,(r, + 10a/i2) S ' h22 t = (L=-l\l a [ 2(C3 + 10a/i2)J The load capacity by integrating (12-82) between 0 and B is given by nr _ k2U2LB2(a + 1) , , , a - 0,JX w = "h^T-TyvW \m + "ln a + !hi) (12'8,}) where m = ———— r + st a -hi s(2at — a + 1) + (a — l)2r n = — a~ — 1
378 Theory of Hydrodynamic Lubrication The equivalent pressure and load capacity equations for laminar flow are from Chap. 3 6/zUB(a - 1) x(l - x) P h22(a + 1) [a — (a — l)x]2 6pULB2 /, 0 a - 1\ w = hxa- irVna-2v+i) The frictional force in the slider is, by integrating the shear stress in the viscous sublayer, * t (B n i &U2L p/1Q , 6C3 . C/\ , Jo ~ 2ir* Jo V h + 2h*) „ kHPBL f„_ . . 36/i2’a + C,1 , y“T»fc1(a-'l)L3C,lna+ 4h2a (a ~ 1}J (12^} with the corresponding drag in the laminar flow given by h2(a - 1) L a + 1 J Because of the original assumption of nd2u/dy2 = 0, the turbulent equations for pressure and load are independent of viscosity. Instead, they depend on fc2, which has the dimensions of density and is an experi¬ mental quantity. Even for low values of k2 the load capacity of turbu¬ lent flow is higher than that of laminar flow. A numerical example for B = 2, U = 2,500, h = 0.002, a = 2, M = 0.145, and Re = 2,410 would yield the following values: k W/L Po H/L 0 2.30 X 10* 1.81 X 10* 0.212 0.01 3.30 X 10* 3.08 X 104 3.93 X 10 0.1 3.30 X 10« 3.08 X 10® 3.93 X 10® 1.0 3.30 X 10® 3.08 X 10® 3.93 X 10® SOURCES 1. Brand, R. S.: Inertia Forces in Lubricating Films, Trans. ASME, vol. 77, p. 363, 1955. 2. Kahlert, W.: Der Einfluss Der Tragheitskrafte bei der hydrodynamischen Schmiermitteltheorie, Ingenr.-Arch., Band XVI, 1948, pp. 321-342. 3. Slezkin, N. A., andJ3. M. Targ: The Generalized Equations of Reynolds, Compl. rend. acad. sci. U.R.S.S., vol. LIV, no. 3, 1946. 4. Osterle, F., and E. Saibel: On the Effect of Lubricant Inertia in Hydrodynamic Lubrication, Z. Angew. Math. u. Phys., vol. 6, p. 334, 1955. 5. Osterle, F., Y. T. Chou, and E. A. Saibel: The Effect of Lubricant Inertia in Journal Bearing Lubrication, ASME Paper 57-APM-37.
Inertia and Turbulence Effects 379 6. Ladanyi, D. J.: Effects of Temporal Tangential Bearing Acceleration on Per¬ formance Characteristics of Slider and Journal Bearings, NACA Tech. Note 1730,1948. 7. Lyman, F. A., and E. A. Saibel: Transient Lubrication of an Accelerated Infinite Slider Bearing, ASME-ASLE Joint Lubrication Conf., Boston, 1960. 8. Taylor, G. I.: Stability of a Viscous Liquid Contained between Two Rotating Cylinders, Phil. Trans. Roy. Soc. London, ser. A, vol. 223, pp. 289-343, 1923. 9. Tao, L. N.: A Theory of Lubrication in Short Journal Bearings with Turbulent Flow, Trans. ASME, vol. 80, November, 1958. 10. Chou, Y. T., and E. Saibel: The Effect of Turbulence on Slider Bearing Lubri¬ cation, Trans. ASME, Ser. E, vol. 81, March, 1959. 11 MlLAOl - CftlTiCA^
CHAPTER 13 NON-NEWTONIAN FLUIDS GENERAL REMARKS The Reynolds equation as originally derived and as applied in the preceding chapters dealt with Newtonian fluids. However, Newtonian fluids are not the only possible lubricants, and non-Newtonian fluids, particularly greases, have their applications. The use of these fluids brings us back to the beginnings of lubrication theory, namely, to the relation between shear stress and strain. To obtain an expression for the study of rheodynamic lubrication, the Reynolds equation has to be reworked to take account of the peculiar viscosity properties of non- Newtonian lubricants. This new form of Reynolds equation should essentially allow for a local variation of viscosity across the oil film. Thus Newtonian fluids having a constant viscosity across the fluid film are only a special case of this more general family of lubricants. From the basic equation of flow dp/dx = dr/dy we have with p a function of x only: T = y (2) + Ci Let us define the flow properties of the lubricant in general form as fy = 'M then by substitution for r and integration U = Jo F [* (to) + Cl] dy + C* (13'1} 9 = fo {fo" F [9 (to) + Cl] dy + C’| dy where C\ and C2 are constants depending on the velocities at the bearing surfaces and q = const at all sections of the film. The results of these integrations depend, of course, on the functional form F(r). 380
Non-Newtonian Fluids 381 If we deal with a Newtonian fluid, the shear stress is constant and F(r) = I = J* n dy ,nd ,,3'2) which is the basis of the Reynolds equation. If the shear stress is (as per example in greases) given by F(r) = - + k then the same expression (13-2) is obtained. However, the effective viscosity governing pressure generation is not the absolute viscosity based on the mean rate of shear but rather is based on the mean slope of the flow curve at the particular section of the film. In general when F(t) is some arbitrary function, neither the viscosity based on the mean rate of shear across the film nor the mean slope of the flow curve is applicable, and Eqs. (13-1) must be properly integrated. BINGHAM PLASTICS (GREASES) AS LUBRICANTS 13-1. Rheodynamic Bearings. Grease is a common lubricant in roll¬ ing-element bearings, and it is also occasionally used in hydrodynamic bearings. The basic properties of a grease can be described in terms of a Bingham plastic. Both Bingham plastics and Newtonian fluids show a linear relation between shear stress and shear rate. The distinguishing feature of the two is that Bingham plastics have a yield value in shear. This value must be exceeded before flow takes place. Therefore, in those parts of the film where the shear stress does not exceed the yield value, stagnant portions of the lubricant, called cores, will be formed. Actually, owing to their soap-sponge structure, greases are both thixo- tropic (the property of coagulating when at rest and of becoming fluid when stressed) and elastic, and various greases approximate a Bingham plastic in varying degrees. Mathematically, the characteristics of greases can be written down as r=±ro + M^ (13-3) dy where r0 is the yield value, a quantity equal to zero in Newtonian flu¬ ids. This relation yields dp/dx = /x d2u/dy2f an equation similar to that of Newtonian hydrodynamics except that this equation holds not throughout the entire rheodynamic film, but only in those regions where du/dy j* 0. By denoting by ha and hb the lower and upper limits of the
382 Theory of Hydrodynamic Lubrication stagnant cores and then considering the equilibrium of an element of the core, we obtain g (h„ - K) = ±2rc (13-4) By applying these equations to a slider bearing with the runner moving at velocity U and the core at some intermediate velocity ue and by inte¬ grating (13-2) for the three separate regions, we obtain “= ~~M{K~v)2 + ^~T^)y~ U 0<y <K (13-5a) u = — uc ha<y<hb (13-56) 1 dp (h + hb y2 hhb\ h — y , . . , M= ~nrx{— y~ 2 -^)-u‘h=hb h^y^h <13-5c> The volume flow across any section must remain constant, thus L \dy = q=- ^g (V + (h- hby\ + h„) (13-6) At the boundaries of the core, at ha and hb, the velocity gradients must be zero. Hence, by differentiating Eqs. (13-5) and equating the derivatives to zero, we obtain: J_ dp _ uc — U 2fi dx ha2 I dp _ Uc 2y. dx (h, — hb)2 (13-7 a) (13-76) By elimination of uc between the two, 1 dp 2 U u dx (h — hb)2 — ha2 and from Eq. (13-4) (13-8) , _ pU(hb ha) /io n\ ±ro " (h-hby - Ka2 (13'9) Equations (13-7) and (13-9) are sufficient to determine ha, hb and uc. The additional parameter here is r0. It follows from Eq. (13-8) that, if \dp/dx\ < 2pU/h2, both ha and hb are unreal and there will be no cores formed. For this condition the velocity gradient is nowhere zero across the film and Eq. (13-6) reduces to = _ _ Eh q 12/i dx 2 the ordinary Reynolds expression. If ha and hb are real, two cases arise. In the inlet region dp/dx < 0, and if t0/m is large enough, then it is pos¬
Non-Newtonian Fluids 383 sible to make the upper boundary of core adhere to the bearing, that is, hb = h. At the outlet edge dp/dx > 0 and there, with a sufficiently low value of to/m, it is possible to obtain ha = 0, and the core will be formed adjacent to the runner. If ha = 0, then ue = U and the flow equation becomes from Eqs. (13-6) and (13-7) q = - | (2* + h„) If hb = h, uc = 0 and then U j, 9- - 3 A. There will thus be three regions in the bearing: an inlet region with hb = h and ha = const; an inter¬ mediate region with no core, and an outlet region with ha = 0 and h = const — 2h. These regions are shown in Fig. 13-1. It is, of course, possible for the values of to/m to be such that the core touches neither the bearing nor runner surfaces. In that case, the cores will be detached and moving within the'lubricant at* the same velocity ue. cases is more complicated. For a plane slider where the film is given by A = fc,[l+(a-l) we have by integrating Eqs. (13-7) with the boundary conditions p( 0) = p(B) = 0 the following relation for the inlet core: i, j, 3a ha — h 2 , 1 -j- a [b) Fig. 13-1. Formation of cores in slider bearings, (o) Detached cores; (6) at¬ tached cores. The treatment of these (13-10) which shows that the extent of the core is proportional to the minimum film thickness. Since ha must always be less than h\ or ah2} a real value of ha is obtained only if a > 2. The condition a > 2 is essentially the requirement for reverse flow in a slider bearing, and this, of course, never occurs with grease lubrication. Thus a bearing with an inlet-outlet film
384 Theory of Hydrodynamic Lubrication thickness less than 2 will always operate without core formation. The condition for complete core formation may be shown to be given by the inequality Toh2 ^ (a - 2) (a + 1) nU ~ a (13-11) The integration of Eqs. (13-6) and (13-7) for the pressure distribution and then for the load capacity yields W = 2 yUB'L (r=Tp(31»2 -§i„? -? + n) <13-12> The frictional drag calculated from the tangential stresses at y = 0 gives for the total drag 2 yUB F — L ^l(2ln2 + Iln|-f + 2i)\ 03-13) Figures 13-2 to 13-4 give these relations in graphical form; they show an increase in both load capacity and frictional drag of the rheodynamic bearing over one with a Newtonian fluid. Detached cores would tend to reduce somewhat the load ca¬ pacity and appreciably reduce the friction, although in all cases they would still be higher than in an ordinary bearing. This increase in load capacity is due to the throttling action of the cores, which essentially produce much flatter and smaller clearances. In journal bearings, as in the case of a slider bearing, cores are formed in the regions of maximum and minimum clearance, as shown in Fig. 13-5. The more or less uniform clearance created by the cores will tend Fig. 13-2 bearings. Load capacity of slider co - C\J k. I t/fi 18 ' 9 -10 * 9 .1 '9 VO 28 9 & /'<6 V A Newtonian fluid 10 Fig. 13-3. Friction in slider bearings.
Non-Newtonian Fluid8 385 to give a pressure profile with a constant, instead of a concave, slope and a higher load capacity. 0 2 4 6 8 10 o * Fig. 13-4. Friction coefficient in slider bearings. Fig. 13-5. Formation of cores in journal bearings. Fig. 13-6. Formation of cores in squeeze films. 13-2. Squeeze Films. Starting with Eq. (13-2), we can write for the sheared and unsheared regions d2u _ 1 dp dy — —r- for Irl > ro fx dx ^ = 0 dy2 (13-14) for Irl < r0 At the core surfaces where |r| = r0 there is no shear and we can define these surfaces by y = y(x)h/2 where 7(3) is symmetrical with respect to y, as shown in Fig. 13-6. We shall consider the squeeze-film effects of a slider moving only normally to
386 Theory of Hydrodynamic Lubrication the bearing surface, and thus the conditions to be met by Eq. (13-14) are those of u = 0 at both surfaces. Also, the velocity and velocity gradients must be continuous along the y = y(x)h/2 surfaces. Considering first the sheared region, we can integrate the first of Eqs. (13-14) by applying the no-slip conditions at y = ±h/2. Then by replacing y by y(x)h/2, we can integrate again for u and Q to obtain <2 = -^£(i-^(2+^) (13-15) The flow can also be obtained from the velocity relation |0.vi dx which when integrated for Q = 0 at x = 0 (since there is no flow there due to symmetry) gives us Q = LVx By equating the two expressions for flow, we have (.-.w+ Considering next the flow in the core to which the second of Eqs. (13-14) applies, we have from dp/dx = dr/dy for r = to at y = y(x)h the following: dp _ _ ro — (—rc) dx yh By eliminating y between the last two expressions, we have h dp/dx 1 _ \2nVx . . to ^ [(A/2r0) (dp/dx)]2 r,h2 U j which is the basic equation governing the pressure distribution. Two limiting cases will be considered: those in which A = 2\kVB/r^h1 is very small or very large. For very small values of A, Eq. (13-16) becomes h dp/dx_ . 2ro For very large values of A} Eq. (13-16) becomes hdp/dxA-* , 0 _ \2yVx ro to/i2
Non-Newtonian Fluids 387 Upon application of the boundary conditions p(0) = p{B/2) = 0, the pressures for the two above equations are 2ro Pa~ o = T (|) - *] (13-17a) - T {[(#) - '] + SP [(I)’ - '•]] The load capacities are WA-> o = Wa-~ = B2Ln 2 h *W3 + W») 4 h (13-18a) (13-186) An examination by graphical integration of the exact Eqs. (13-16) and (13-18) showed that the WA-*» solution is almost exact for values of A > 3. In ordinary bearings the value of 2fiVB/roh2 is very unlikely to be below 3, so that Eq. (13-186) can be considered the exact solution. Figure 13-7 shows a comparison of load capacity for the two lubricants, a grease and a Newtonian fluid of the same mobility. 14 12 10 <*! 8 / / '/ Bingt iom pi lostic ~y A '/ / V / Newt onion fluid Y I 0 2 4 6 8 (0 12 ZVB\kh„hl Fig. 13-7. Load capacity in squeeze films. 13-3. Rheostatic Bearings. Consistent with the term “rheodynam- ics” as applied to hydrodynamic bearings, rheostatics refers to hydro¬ static bearings operating with grease as a lubricant. The configuration considered is the simple thrust bearing shown in Fig. 13-8. The analysis will consist of two parts, the first a simplified version in which it is assumed that the pressures developed depend only on the radial velocity, although in actuality the lubricant, owing to runner rotation, flows in a spiral. This assumption holds as long as the stresses due to rotation are
388 Theory of Hydrodynamic Lubrication small compared to those of radial flow. Subsequently we shall deal with the dynamics of flow considering both the radial and the circumferential flow components. The boundary conditions for this problem are u(h/2) = u( — h/2) = 0 and continuity of both u and du/dy at points where r = r0. By denoting by y(r)h/2 the boundaries of the core, i.e., where r = r0, we have by Eqs. (13-15) and (13-16) 0--T5-1;! <’•-* + *> and since dp/dr — —2T0/yh, we have after eliminating y (r0 df) + 3 “ [(A/2t,)(dp/dr)]* = ~3A "F {13"19) where A = irh2ToR2 Equation (13-19) is a nonlinear differential equation, and in order to integrate it, some sort of simplification in — [(h/2r ){dp/dr)]2 1S nee(*e(** By setting ( h dp\ \2ro Tr)RlA ~ ^ we see that if the slope at Ri is taken, the resultant pressures will bound the exact pressures from above; if R2 is used, the resultant pressures will constitute a lower bound. With £1,2 a constant Eq. (13-19) can be integrated to yield Ph _ “ I . . I A 1« ‘ I - I , ■ 1 /,o 20) 2t qR2 2[(1 **) "41n«J 281.1* {l «2) (13'" the constant of integration Q, as determined by the amount of flow supplied, being contained in A. By writing p = R\/R2 and using P — W/tR22, we have by integrating Eq. (13-20) Equation (13-21) contains the constant A, which depends on Q. Si,2 can be evaluated in terms of A from Eq. (13-19) by setting either r = Ri or r = R2 and thereby yield
Xon-Newtonian Fluids 389 Thus, for a selected flow Q, A and then Si and S2 can be calculated and used in Eq. (13-21) to supply an upper and lower bound to the load capacity of the bearing. If in Eq. (13-19) to is set equal to zero, this equation integrates to the load capacity of a hydrostatic bearing operating with a Newtonian fluid, namely Ph = 3Qm(1 ~ P2) ich2 Figure 13-9 shows the load ca¬ pacity of the rheostatic bearing, which is seen to exceed that of a plain hydrostatic bearing. It is interesting to note that a rheostatic bearing is able to support loads without actual lubricant flow. 1.2 Fig. 13-9. Load capacity of rheostatic thrust bearing with radial flow only. If the pressures in the pocket do not exceed the yield value, then no flow will take place. Mathematically this implies A = 0 and Si = S2 = — 1. Equation (13-21) then yields Poh TqR2 2(1 - p3) (13-22) where Po is the unit loading that the bearing will support with a grease whose yield value is to. The frictional torque for a rheostatic bearing can readily be calculated from du _ ro) dy ~ h The tangential shear stress thus is 1 ro) T =r» + MT /Rt fRt t(2tct dr)r = 2tttq I This integrates to r2 dr + p I r3 dr P TO M 2r tR2o , . = tt + m or— (1 - P4) R23 3 ' " 2/ito (13-23) where (2tt/3)(t0/?23) is the starting torque.
390 Theory of Hydrodynamic Lubrication In considering both radial and circumferential flow, the first question that must be answered is what combination of stresses is required to cause the grease to flow. This question is not yet adequately settled. Here the yield value will be taken to be the root mean square of the maximum shear stresses acting on three mutually perpendicular planes. This quantity, the invariant of the reduced stresses, gives in cylindrical coordinates where the Sij terms are the reduced stresses and the a terms are the fluid stresses. Flow will then occur only if J2 exceeds a certain critical value. the condition that the ratios of deformation rates must equal the ratios of the reduced stresses. By denoting by the components of deforma¬ tion rate, we have which is a generalized form of the relation given at the beginning of the j 2 = H(«rr2 + See2 + S«2) + Or# + O>02 + VtO2 The formulation of a stress-deformation rate relation must also satisfy 1 y/ J i — ro y/J2 > To V J2 < To d.. = /2m y/jt atJ — V X0 (13-24) chapter for Bingham plastics. The stress-strain equations in cylindrical coordinates are [%<Trr ~ %((T0e + V,z)] [%Oee — M(*rr + O] M(°Vr + <Jee)] (13-25) The sum of the first three equations gives the continuity equation
Non-Newtonian Fluids 391 The force equations with the three inertia terms neglected yield d<rrr 1 d<rre , do- rz , &rr 06$ _ q dr ■*" r dd dz ^ r d^ + -dj£+ *-¥ + — = 0 (13-27) dr r dd dz r d<Trx . 1^ d<Tx$ . dozx . (Trt ~ ~dr rdd dz r — A solution of the problem of Fig. 13-8 has to satisfy Eqs. (13-26) and (13-27). Proceeding from the physical fact that the flow of the lubricant is determined primarily by the rotation of the runner, we shall assume that Ji is influenced mainly by the circumferential shear stress. The problem is thus solved for rotation alone. Then to the obtained veloci¬ ties and pressures we add the contribution resulting from radial flow. This will be done by perturbation methods. Considering the tangential flow only, we have for the configuration of Fig. 13-8 z . rco u = ru - <jte = r0 + M -jr v = w = 0 srr = See = szt = <r„ = <rre = 0 By adding the radial flow and expressing the above terms in somewhat different form, we have Vibiz 2 , V mi. V max v = -5— Vl u = ro) -r + u 1 w = Wi K20) fo ti 200 K203 VtDMX V m»x Vmtx $rr = "n ^rr 1 S$e = S&81 Szz ~ $zzl it 2W it 2(i) it Vm*x I r(t) . i'max Vmmx <Tr6 ~ d— <7rei ^9 “ TO + M T <?z61 T* 0x61 Ort — -5— Orzl lt2W fo Xl2C*) where Vm*, is the maximum radial velocity and subscript 1 denotes the perturbations on the corresponding quantities. In all these equations By using the last expressions in the equation for the invariant J2, we have J2 = (T» + * t) + (to + mt) which gives Eq. (13-24) the form w.. - J-V2i To 1 I” 1 I MmxxTo&zlH "J ..n OQ\ a<’ = %. -JIT ~ ^ L l+T*h/nro, + 7e2«(ro + ,mo//t)2J l
392 Theory of Hydrodynamic Lubrication Assuming that Ui, that is, the contribution of the radial flow to the circumferential velocity, is zero, we have, by using Eq. (13-28) in Eqs. (13-25) to (13-27), the following: IF = i (r+ r[%a"' ~ 1AW,n + 7 = i (l +r!fc/„r«) {%atn ~ H(arr' + °",)l 1 = % (rrrkvi)[%a-1 ~ MKrl + <79,,)1 i dm = 1 / 1 \ dz dr u \1 + Toh/yru/ zrl <r,e i = 0 0*«i = 0 Id, . , dwi n - 3- {rvi) + — = 0 r dr dz dflVrl dffui fffd o~$$i q dr dz r du> dz dy (13-29) dcrrti + — = 0 dr dz r By eliminating the unknowns tt>i, <rrri, <rm, <r«i, and <r„i, we obtain the following equation for t>i: d4i>i 0 d4^i d4vi 2 d*vi 2 d3t>i 3 d*i?i 2 d2i>i 3 d^ dz4 + dr2 dz2 + dr4 + r dr3 + r dr dz2 r2 dr2 r2 dz2 r3 dr 1 d2i>i 5 diij 5 \ , The homogeneous solution of Eq. (13-30) satisfies the boundary con¬ dition Di = 0 at z = ± h/2, so we write for the particular solution the expression . Toh H fxroo c, where Ci is a constant and t>n is a function to be determined. By intro¬ ducing this expression for Vi into Eq. (13-30) and retaining only first-order terms of Toh/uru (roh/yro) <K 1), we get d^Vii , 2 d4Vn , d4*>n , 2 dh>u , 2 d3yn _ _3 d2vn _ 2^ d2vn dz4 dr2 dz2 dr4 r dr3 r dr dz2 r2 dr2 r2 dz2 , 3 dvn 3 _ 2C\ + ? d7 “ 74 Pn " 7^
Non-Newtonian Fluids 393 By representing tin by a power series of the form «o = 2 /*«»* *- -2 and remembering that, since z is here in the direction of the film thickness (z/r 1), we get + °lfzi > n i . lA 16C» ( z* . Dzi . n - f(I2 + *** + *)- — (360 + 12 +^*+c) and for t>i Pl_£!(,._V/4)+^§[(g + &* + £) _S(sio + :S"f^ + G)+ ‘ ' '] From the condition V\ = 0 at z = ±h/2 we reduce the number of con¬ stants by two, which yields (<•-1)-?)+•] t’l r I 16 360r2 The condition Wi = 0 at z = 0 due to symmetry and wx = 0 at ±h/2 can be used in the continuity equation Id. . dwi r dr ~dz = to obtain D and F which become D = — — F = 9h* 40 16 • 40 • 35 When h/r terms of order higher than one are dropped, the expression for Vi becomes (13-31) The remaining constant Ci can be expressed in terms of flow into the bearing. Since the integrated bearing flow must equal the input flow, we have r h/2 v fh/2 Q - 4r / rvdz = 4tt rvx dz J0 iV2« Jo
394 Theory of Hydrodynamic Lubrication which yields p _ 3 Q R20) 1 7r/l3 i>m»x Finally ■ra[I_ 4(*)’]) “”2) From the first three equations of set (13-29) and by using the boundary conditions for the pressure (Trr = 7>o at r = R2 we obtain for the radial pressure distribution (1M3> The pressure at which the lubricant must be admitted is obtained simply by setting r = R\, which gives us P.. = Pa + ^ [hi ^ (|i - 1)] (13-34) where pa is the atmospheric pressure and p,a the absolute supply pressure. The total load capacity is given by W = — f * azi(2irr) dr — TRi2<rtt — ttR22<tTt — — 2irf pr dr JR1 r««i JR1 — ir(Ri2pta — Ri2pa) which upon integration gives 6nQ IR22 — Ri W = yR2o) |_ V*V 2 \ti2/ 2jj k3 \ 2 The torque required is M = 2tt {R,<r,<r*dr J R1 which gives -^[iO-wO + ireO-®)] These equations rewritten yield W (ir/2)[(l - RS/RS) - (2Tgh/pRiw) (2flt2/R22 - dRj/Rj + 1)] ln (R2/R1) + (rJi/nR#»XRt/Ri — 1) (13-35)
Non-Newtonian Fluids 395 3.0 2.5 2.0 ‘ 1.0 0.5 0 0.02 0.04 0.06 0.08 r0/f/2fiF2U) 0.1 20 18 Wi 16 = 0.8 14 £1?'2 = 0.6 Ml" 10 ti*8 = 0.4 6 4 =0.2 0 = 0.01 C 0, Fio. 13-10. Load capacity for rheostatic thrust bearings. where p, is the gauge inlet pressure. Mh 0.04 0.06 0.08 T0/t/2fiR2(o Fig. 13-11. Pumping-power requirements for rheostatic thrust bearings. For the torque 2mo)R24 t [4 (X R2*) + 3 2m«2«(1 /f23)j (13'36) and the power required to feed the lubricant is 5W-T['4l + ^;(t-■)]"' {,M7) These equations reduce to those of ordinary hydrostatic bearings when to is set equal to zero. Figures 13-10 to 13-12 are plots of the last three equations. Contrary to the conclusions reached for the bearing with radial flow only, the load capacity of grease is less than that of Newtonian fluids. The higher the value of ro, the lower the load capacity. The frictional drag is, as before, higher for a rheo¬ static bearing and is almost a linear function of r0. The assumption made in the analy¬ sis to the effect that v^/R^w <3C 1 imposes the qualification that the circumferential stresses be high com¬ pared to all other remaining stresses. By a review of the order of magnitude of the stresses and reduced stresses involved, it can be shown that this condition implies Fig. 13-12. Torque requirements for rheostatic thrust bearings. n 9 In ^2Y p‘ W~lnRi)
396 Theory of Hydrodynamic Lubrication Since h is usually very small and w high, the condition is true in most rotating bearings. VISCOELASTIC LUBRICANTS A viscoelastic fluid is one that exhibits properties of both a viscous fluid and an elastic body. All liquids are to a degree viscoelastic; the prominence of the elastic behavior depends on the rate of shear. If a liquid is stressed rapidly enough, it will show an elastic response to stress, and thus the duration of stress becomes a parameter in describing the properties of such fluids. Assuming linear behavior, we can say that the total distortion of an element of fluid is the sum of the elastic deformation and viscous flow or 1 1 5 + r- (i M where G is the shear modulus and r the applied stress. The last term in the above equation is the rate of flow, and we must think in terms of rate of flow because its value depends on the duration of stress r. Thus we must talk of rates of deformation, and the above equation should be written as ar-r + si (,3‘38) If an element of fluid is rapidly deformed and then constrained in its deformed shape, the internal flow gradually relaxes the stress. Under these conditions we eventually have db/dt = 0 and Eq. (13-38) becomes 1— 4- I - n G dt + „ T which integrates to r = roe-^ (13-39) The last expression tells us that the stress relaxes exponentially with time. After a time interval equal to n/G, r will have decayed to (l/e)r0, where r0 is the original stress. \l/G is called the “relaxation time.” If the stress acts over a time interval T, then the deformation is given by Elastic deformation = ^ Flow deformation = - T The two terms can be compared to each other as
Non-Newtonian Fluids 397 We thus have If T » £ If T « £ Newtonian fluid elastic body If T = 0 viscoelastic fluid If n/G becomes of the same order of magnitude as the time during which the lubricant is being stressed (duration of flow through loaded part of bearing), then the lubricant will act as a spring. Another important iy factor from the standpoint of lubri¬ cation is the so-called Weissenberg effect. This is the phenomenon of producing a force normal to the i-k,. direction of shear, which will act J normal to and along the journal. Both this tendency and the spring ^ ► action of the fluid may influence u appreciably the operation of bear- Fig. 13-13. Bearing configuration, ings and gears at high speeds. To get an insight into the ramifications of using viscoelastic lubricants, let us consider two nonparallel surfaces, both moving with a tangential velocity U as shown in Fig. 13-13. We shall for the sake of simplifying the solutions assume an exponential film shape h = he~ax, and we shall also assume the following: 1. The usual assumptions of hydrodynamic lubrication. 2. The relation between stress and strain, taking into account some of the previous assumptions that v, uf and h are very small, is by Eq. 3. The pressure terms due to the local elastic components of strain are assumed to be small compared with the ordinary hydrodynamic forces, an assumption that may not hold at very high rates of shear. Under the assumed conditions we then write (13-38) (13-40)
398 Theory of Hydrodynamic Lubrication Since for the configuration of Fig. 13-13 the inlet-outlet film thickness is not much larger than one and both surfaces move with a velocity U, it can be assumed that the velocity of the fluid dx/dt = U and thus <iwi> Upon substitution for r of its value in terms of dp/dx du dy integration 1 ( dV , ^ \ , U ( dtp , dC A = »(ydx + c')+G\yd£+*;) u = yl(i^ + lL^\ + v(c1+udc\ 2 dx + G dx1) + V Vm G dx) ^ and the use of boundary conditions u = (0) = u(h) = U, the expression for u becomes -"+i<«-*>ei+ss) <>«*> By integrating for q = fu dy, we obtain JJh hz /I dp U d2p\ , . . 9= Uh ~ ni^Tx + a w) (13^J) or as a differential equation S If l/G in the above expression is set equal to zero, the equation reduces to the familiar one-dimensional Reynolds equation. The integration of Eq. (13-44) with h = h\e~ax is straightforward and yields - r Mr .-fix l2W $e*az m 12^u Pe2ax p(x) Cj + C,e 3a(3a + ^ 2a(2a + ^ (13-45) where @B = GB/Uy is a measure of the ratio of the transit time B/U to the relaxation time y/G and Cz and C< are the constants of integration. p dx gives w - l [c,B + £ (, - «-») - “a (13-46 The frictional drag on the plane y = 0 is given by F = J — Ijt | Q dx or after proper integration F = L [— (1 — e~pB) - l2-q ^22 ~ ^ ~ ^ 1 (13-47) [ 0 hi2 4“(2<* + 0) hi 2*(<* + 0) J
Non-Newtonian Fluids 399 which also contains two arbitrary constants Ci and q. The two constants in Eq. (13-46) can be determined from the boundary conditions p(0) and p(B), but the other two depend on the state of the fluid at the moment of entry. Physically it means that the moving planes subject the lubricant to shear even before it enters the clearance space. If the shearing is slow as compared to the relaxation time, the lubricant will enter the bear¬ ing in a partially relaxed state. If the shear is rapid, the lubricant will enter the bearing unrelaxed and receive a suddenly applied rate of shear. The two extremes of lubricant entry are: 1. Fully relaxed, that is, r/y = du/dy, which would be applicable to Newtonian fluids 2. Unstressed, that is, r — 0, which is more applicable to cases where the transit time B/U is comparable to the relaxation time The actual state of the fluid probably lies somewhere between the two extremes. Considering case 1 first, we have t du - = — at x — 0 dy This from r = y ^ + Ci gives du 1 / dp n \ dy y V dx 7 which when used in the equation following Eq. (13-44) gives U( d2p,dC1\ _ n G y dx2 dx ) or since this applies to all the fluid at the inlet tE = = n dx2 dx u The full set of four boundary conditions is thus given by p = S = 7E = 0 at x = 0 p = 0 at x = B which when used in Eqs. (13-44) to (13-47) yields 12m U 0 Cz = W 2a(2a + 0) (a2 - 1)(02 - 9a2) - (a3 - 1)(02 - 4a2) - 5a2(l 02(a3 - 1) + 9a2(l - e-»B) 12MU af} 9(a2 - 1) - 4(as - 1) hi2 2 (2a + 0) 02(a3 - 1) + 9a2(1 - e~»B) - 3 Tlh 3a + 0 02(a2 - 1) + 4a2(1 - <r»*) 9 2 1 2a + 0 02(a3 - 1) + 9a2(l - e~BB) „ 12^9 ot 12 uU Cb — » o• ^ 1 hx2 2ot (5 hi 2(a + 0)
400 Theory of IIydrodynamic Lubrication These values inserted in Eqs. (13-46) and (13-47) yield w ^ nULB2 16&i ln a[(a2 - 1 )(kl2 - 9) - (a3 - l)^2 - 4) - 5fc2] (13-48) (13-49) h22 \ a2 (In2 a) (2 + IfciP^a* - 1) + 9k2] k2[27(a2 - 1)(2 + fcO - 8(a3 - 1)(3 + kx)] + fc^a2 - l)(a3 - 1) “l" a2 (In2 a)(2 + fci)[/ci2(a3 - 1) + 9A;2] _ 3/it/LB (4(2 + ki)2[ki2(a* - 1) + 9k2][kx2(a - 1) - k2] 2h2 { a(ln a)fci(l + kx)(2 + fci)2[fci2(a3 - 1) + 9/c2] 3(3 + *0(1 + ki)[k\2(a2 - 1) - ^k2][kx2(a2 - 1) + 4fc2] a(ln a)fci(l + kx)(2 + fcOWfa* - 1) + 0/c2] where k\ = - — ^ and k2 = I — e~&B a Ujxa If the lubricant is assumed to be unstressed outside the bearing, then by writing r = 0 at x = 0 ,i+c..o or since this condition applies at any y Tx + Cl = 0 and the four boundary conditions are p = g^ = Ci = 0 at x = 0 p = 0 at x = B which gives for the values of the integration constants 0 3 ft,2 2a(2a + 0) (a2 - l)(3a + 0) - (a3 - l)(2a + 0) + a(l - e~fB) 0(a3 - 1) - 3a(l - <r«B) „ YZnU 0 3(a2 - 1) - 2(a3 - 1) t 4 = — ft,2 2(2a + 0) 0(a3 - 1) - 3a(l - «"«) , = 3 ... 3a + 00(a2 - 1) - 2a(l - e-W) q 2 1 2a + 0 0(o3 - 1) - 3a(l - e~BB) r _ 12nU 0 129m 0 ^ ft, 2(a + 0) ft,2 2(2a + 0) The load capacity and friction are then given by |6ft, ^ WAIn* _ nULB2 (6ft, ln a((o2 - 1)(3 + ft,) - (a3 - 1)(2 + ft,) + ft.) ft22 \ a2(ln2 a)(2 + fc,)[fc,(a3 - 1) - 3fc2] 2[4(o3 - 1)(3 + ft,) - 9(a2 - 1)(2 + ft,)) + fc,2(a2 - l)(a3 - 1)1 a2(ln2 a)(2 + fc,)[ft,(o3 - 1) - 3fc2] j (13-50)
X on-Newtonian Fluids 401 F = Z»ULB 4(2 + kiyikfa* - 1) - 3fc2Pi(<* - 1) + k2) 2h2 [ a(ln a)( 1 + *i)(2 + *i)2[*i(«8 - 1) - 3k2) 3(3 + fei)(l + ki)[ki(a2 - 1) + 2fe,p1(o« - 1) - 2k2]\ a(ln a)(l + ki)(2 + ki)2[ki(a* — 1) — 3A;2] j (13-51) In all expressions above the ratio pB = (B/U)/(n/G) is the ratio of transit time to the relaxation time. When pBy> 1, the equations for W and F reduce to the ordinary equations of hydrodynamic lubrication: lim SB-* * lim SB-* oo W = nULB* h22 F = nULB In2 3 h2 2a ln <‘«2> -4^] A comparison of load capacity and friction values for lubrication with Newtonian and viscoelastic fluids is given in Figs. 13-14 and 13-15. _ 0.4 *5 0.3 •>0.2 ^0.1 % Newtonio n fluid 1 I y* V 1 0.5 10 100 500 Rotio of transit time to reloxotion time Fio. 13-14. Load capacity of bearings using viscoelastic lubricants; a = 2. (A) The lubricant is assumed to enter the bearing in an unstressed condition. (B) The lubricant is assumed to be fully relaxed at the moment of entry. 1.2 1.0 ^ 0.8 |'0.6 S~0.4 0.2 V - Xs Newtonior i fluid \ 0.5 10 100 500 Ratio of transit time to relaxation time Fia. 13-15. Friction in bearings using viscoelastic lubricants; a = 2. (j4) The lubricant is assumed to enter the bearing in an unstressed condition. (B) The lubricant is assumed to be fully relaxed at the moment of entry. It can be seen when @B is low, say, up to 10, the load capacity is much lower and the drag much higher for a viscoelastic fluid. Since cases 1 and 2 are the two extremes and the actual fluids are somewhat between the two, it follows that the elasticity of a lubricant decreases sharply the load capacity of bearings. SOURCES 1. Milne, A. A.: A Theory of Grease Lubrication of a Slider Bearing, Proc. Second Intern. Congr. of Rheology, 1958. 2. Milne, A. A.: A Theory of Rheodynamic Lubrication, Kolloid Z.y Band 139, Heft 1/2, p. 6, 1954.
402 Theory of Hydrodynamic Lubrication 3. Osterle, F., A. Charnes, and F. Saibel: Rheodynamic Squeeze Film, Lubrication Eng., vol. 12, no. 1, pp. 33-36, 1958. 4. Osterle, F., and E. Saibel: The Rheostatic Thrust Bearing, ASME Paper 55-LUB-6. 5. Slibar, A., and P. R. Paslay: On the Theory of Grease Lubricated Thrust Bear¬ ings, ASME Paper 56-LUB-l. 6. Milne, A. A.: Theory of Rheodynamic Lubrication for a Maxwell Liquid, Conf. on Lubrication and Wear, Paper 41, London, 1957.
CHAPTER 14 EXTENSION OF THE CLASSICAL THEORY THE RESTRICTIONS OF LUBRICATION THEORY The Reynolds equation and the theory based on it represent the basic mechanism of hydrodynamic lubrication. The equation rests on the various assumptions stated in Chap. 1, most of which stem from the physical fact that the height of the fluid film is very small compared with the axial and longitudinal dimensions of the film. As a conse¬ quence, it is the practice to neglect all hydrodynamic variations across the fluid film. It is also the practice of conventional analysis to confine the region of investigation to the actual bearing clearance, since, as seen above, the Reynolds equation is valid only for such thin films. It is usually assumed that at the edges of the bearing, the pressures are the static values of the surrounding atmosphere, and the hydrodynamics of the problem are confined strictly to the interface between the two mating surfaces. Both of the above assumptions will now be removed. We shall try to generalize the Reynolds equation by allowing the exist¬ ence of fluid films of any arbitrary thickness. We shall also, wherever applicable, extend our field of interest and try to take into account any hydrodynamic effects in the fluid surrounding the actual bearing. This extension of the classical theory should accomplish several things. In the first place it will extend the applicability of our equations to such devices as sliders with high tapers, cylinders with relatively high clear¬ ances, piston rings, electric brushes, and many other similar configura¬ tions. Secondly, this may help explain one of the puzzling phenomena of hydrodynamic lubrication, namely, the successful operation of parallel sliders and centrally pivoted pads, both of which should by conventional theory (neglecting pad deformation and thermal effects) have no load capacity. By removing the restriction of a very thin film and by con¬ sidering the hydrodynamics of the fluid at the bearing ends, we can extend the validity of the hydrodynamic equations to both parallel and highly divergent fluid films, the two extremes at which the classical Reynolds equation collapses and fails to provide solutions. Even for infinitely long bearings, the problem becomes two-dimen¬ sional; for all quantities will now vary along and across the fluid film. 403
404 Theory of Hydrodynamic Lubrication The equations to be solved are the two basic equations dp _ d/u dp _ d/v dx M dy2 dy M dy2 Let us introduce a stream function ^ such that d\p dyp u = ~ v = — -r1- dy dx By using stream functions in the preceding equations, differentiating dp/dx with respect to y and dp/dy with respect to x, and subtracting, we have the biharmonic equation . VV = 0 (14-1) The hydrodynamic problem thus becomes one of finding a function \p satisfying Eq. (14-1) and matching the appropriate boundary condi¬ tions. Solutions to this equation under various conditions are given in subsequent paragraphs. THE INFLOW WEDGE 14-1. The Case of Parallel Sliders. We shall, by employing the sim¬ plest possible approach, attempt to show that a slider moving submerged in a fluid can build up hydro- dynamic pressures in front of the leading edge, much as stagnant pressures are produced in pitot tubes, ram jets, and similar arrange¬ ments. In this case the parallel film with the inflow pressures be¬ comes the equivalent of a composite bearing discussed in Sec. 3-10, which of course is capable of sup¬ porting a load. The only difference is that the wedge instead of being formed by and contained within the bearing geometry is now formed in the inflow region, in front of the leading edge of the parallel slider. Let us consider the dynamics of the inflow region yOx as represented in Fig. 14-1, where the leading face yO of the slider is pressed flush against the runner Ox moving with a velocity U so that no lubricant is admitted to the interface, or h — 0. The bodies are assumed to be rigid and infinite in extent, and the lubricant is assumed to be isoviscous and incompressible. Inertia forces are neglected. Fig. 14-1. Fluid element in the inflow wedge.
Extension of the Classical Theory 405 Because of the geometry of the corner yOx, a certain symmetry about the origin 0 is expected. For a first approximation let it be assumed that the flow is purely radial, and we shall call this approach the polar approximation. It follows immediately that, since there is no flow in the angular direction, all 0 = const lines are also isobars. It further follows that there are no tangential stresses in the 0 direction and that the force equilibrium equation is simply dp d2u ( . dr - *1? (14'2) where ds = r dd. Since for a constant radius the pressure is independent of 0, it follows that it is also independent of s. Thus for r = const, dp/dr is also independent of s and we can integrate Eq. (14-2) along a constant r. By using the boundary conditions u — U at s = 0 u = 0 at s = H = rfi with 0 the angle between runner and front of slider, we obtain »-<«-•) (it+») <■«> The total flow across any surface r = const must be zero, since there is no loss of fluid past the origin, and thus [” , Hz dp « = j0 uds = 12^37 Hence dp _ 6mU _ 6yU df ~ ~ w* (14-4) By using the condition p = 0 at r = oo, we can integrate Eq. (14-4) for the pressure distribution P(r) = ^ (14-5) By putting the expression for dp/dr into Eq. (14-3), we obtain for /3 = tt/2 It is seen that u = 0 at ?r/2, which is the boundary, and also at 0 = ir/6, which indicates a reversal of flow in this plane as shown in Fig. 14-2. from this last equation occurs at 7r/3, and its value is 17/3. In general, for any value of P we have u = 0 at 0/3 um„ = - at 20/3
406 Theory of Hydrodynamic Lubrication Equation (14-5) gives positive pressures even for (3 — tc, which is the movement of a flat plate in an infinite fluid. The pressure build-up is then one-quarter that obtained for a right angle. The effect of a pressure build-up in this inflow region is essentially to im¬ pose a positive boundary pressure at the leading edge of^the slider. Thus, for a slider with a clearance space of a uni¬ form thickness h we have dp _ dx M dx2 d2u _ pc ~ B (14-6) where pe is the pressure of the leading edge of the slider to be determined. That the value of dp/dx is a con¬ stant is clear from the fact that the streamlines are parallel to the interface h. Equation (14-6) with its proper boundary conditions for velocity integrates to Fig. 14-2. Velocity field in the inflow wedge. and fh , ft'p. , UH ~q = Jo udy = WB + ^~ (14-7) (14-8) The flow in the inflow region with h no longer equal to zero is by the approximate formula above IJH 2 Hz (dp\ , UI q I2n \dr) + 2 By solving for dp/dr, integrating, and assuming p = 0 at H = oo, we obtain for a square slider (14-9) By using for q the value from Eq. (14-8) and setting p = pe when H = h, we have 6m U Pe (ir + h/B)h The load capacity is then simply /> = £?= :*Mf 2 Or + k/Ii)h irh (14-10) (14-11)
Extension of the Classical Theory 407 For an arbitrary angle the relation between h and P is The stress near the moving surface is du | yUv — 2h/B yU ,1A 10X T=-^U = T7+W“T (14'13) The total drag is then F = L t dx = BLt Hence the coefficient of friction is in general “4-M> By a similar treatment on the trailing side it is easy to show that the trailing boundary pressure is given by Pi ~ (i^b)2 p‘ (14'15) which is a small value. The above problem can also be solved without the restricting assump¬ tion of polar symmetry. This can be done by superimposing the solu¬ tions to the following two problems: (1) Find the pressure distribution in a sector-shaped inflow zone with one moving boundary and no leakage. (2) Find the pressures in a similar sector with all boundaries stationary but with a point sink of strength q = Uh/2 at the apex. The two stream functions satisfying the biharmonic equation VV = 0 for cases 1 and 2 are Ur = 02 _ sin2 p I9 sin P sin (0 ” e) “ 0(P ” 6) sin 0] (14-16a) _ 2g[sin (0 - 26) + 26 cos 0] , +* ~ 2 cos 0(2/3 - sin 2/3) - sin* 0 (14'16b) From m = - and v = — ~ r dd dr Ul = 42 _ dni “ fS‘n ^Sin (P ~ 0) ~ 6 COS (0 - 0)] U_ p2 — sin2 P + p[sin 0 — (p — 8) cos 0]} (14-17a) «i = 02 _^in; ^ (6 sin 0 sin (0 - 0) - 0(0 - 6) sin 6} (14-18a) 8q ( sin 6 sin (0 — 6) ) r (sin30 - 2 cos 0(20 - sin 20)) (14-176) vt = 0 (14-186)
408 Theory of Hydrodynamic Lubrication When Eqs. (14-16) are used in Eq. (14-3) and integrated, we have = 02 -^2 0 [0 sin 0 — sin /? sin (/3 - 6)] (14-19a) _ 0 sin 20 + sin 2(/3 - 0) . iqm nq 1 - cos 20 - 0 sin 2/3 1 ; For 0 = 1 ~f[j ~ IL 4 ^ s^n ® — cos (14-20a) = -2 sin 20 (14-206) nq Equations (14-19) and (14-20) are now seen to be dependent on both 6 and r. pi gives values that are generally lower than those given by Eq. (14-5), although the maximum pi which occurs at 0 = 57J^° is somewhat higher than the value given by Eq. (14-5). The points at which Ui = 0 occur at 31° for /3 = ir/2 and at 64° for 0 = tt. The corresponding values for the polar approximation are 30 and 60°. To get the load capacity of a parallel film, we must now add pi and p2 at the entrance. This can be done in either of two ways. By setting directly 6 = t/2 and r = h, we obtain from Eqs. (14-20) and so for a linear pressure distribution in the film P = y2Vc and h = 1.07 ^ (14-22) If, however, we average pi and p2 over 0 (Pi A = 8 / Pir2\ = _ 4 \M U/»vt *(* - 2) \ nq /avg 7T P.v. = (p. + p2)...=^(^f-f) For r = h and q = Uh/2 we have 4/2 1 \y.U Ve ir\x - 2 2/ h (14-2:1) The load capacity then is = h = n 7Q P = VzVc h = 0.79 (14-24) The values of h obtained from both Eq. (14-22) and Eq. (14-24) are seen to be not too far from the value of 0.955y.U/P given by Eq. (14-11).
Extension of the Classical Theory 409 The results contain two significant features: First, because of the linear pressure distribution over the slider from pe at the leading edge down to near zero at the trailing edge, the resultant load acts through a line located at x = B/3 from the leading edge. This means that an external load applied in the center of the slider would cause the slider to tilt and thus produce a hydrodynamic wedge. This may throw some light on the operation of lightly loaded or very rigid centrally pivoted tilting pad bearings and on the success with parallel plate sliders in general. The other novel result is the nature of the coefficient of friction, which for a square slider is nearly t/3, or greater than unity. This is a very high value, and it suggests that high friction is not always an indication of the onset of boundary lubrication. It is also noteworthy that the coefficient of friction is independent of load, speed, or viscosity. Fig. 14-3. Experimental streamlines in Fig. 14-4. Inclined slider with inflow the inflow wedge. wedge. Experiments1* yielded the flow pattern given in Fig. 14-3. The flow of lubricant into the clearance space essentially came from the region underneath the ^ = 0 line. This streamline starts at the tip of the slider and tends asymptotically toward the runner surface. It is highly pressurized, and it may thus be considered an extension of the hydrodynamic film. The u = 0 line too originates at the tip of the slider and then tends asymptotically toward the 6 = 30° line measured from the origin (0,0) and merging with the tt/6 radius at a distance of about 10h from the origin. The liquid contained between ^ = 0 and u = 0 lines begins by moving inwardly, crosses the u = 0 line, and then flows outward in the region above u = 0. Above u = 0 all streamlines issue from the leading tip of the slider; below u = 0 they are an extension of the streamlines in the film. 14-2. The General Case of Plane Sliders. In the experiments men¬ tioned above the flow pattern was observed to remain the same even when the sliding was not parallel. Thus the treatment employed above can be extended to include inclined sliders. By referring to Fig. 14-4 and writ- * Such superscript figures indicate references listed under Sources at the end of the chapter.
410 Theory of Hydrodynamic Lubrication ing a = h\/h2, where a may now also assume the value of unity, we can start out with Eq. (14-3) for the inflow region and write « = + - _ A37*3 9 “ 12/i dr ~2~ The pressure in the clearance space is no longer linear and the equivalent two equations for the interface are = _ _ Mh 9 12/i dx 2 From continuity we must have Uh q = —= const By using this last value in the inflow region with the boundary condition p = 0 at r = *, we have P‘ = ^ShX' ~ 2oii) (14-256) By using q = Uh0/2 in the clearance space with the boundary condition p = 0 at h = h2 3*UB (h - h2)[(h + h2)h0 - 2/i] p ~ (T-~Dh? ~ P (14'2b) = (\ _ m 97i Pc a0ht\ 2 a) h$[a + 1 + (kt/0B)] , ho 2a(l + ho/0B) where h„ = -^ = —- , t ,-' h2 a 1 ”f" h2/(3B The maximum pressure occurs at h = ho = A 2^0, or 3/.t/fl (Ao-1)2 p° = -T— (14‘28) Several pressure distribution curves are given in Fig. 14-5. It can be seen that it is possible to have p0 occur at the leading edge of the slider. This occurs when h2/R = (a — 1)0. Since 0 is a value close to tt/2, it
Extension of the Classical Theory 411 follows that the relation is valid only for values of a close to unity or for fluid films of very small convergence. By replacing the values of. dp/dr in the expressions for u, the velocities in the clearance and in the inflow x/B Fio. 14-5. Sample pressure distribution. Fig. 14-0. Sample velocity distribution, region are given respectively by “-(‘-!)[t(1-i)-1]'/ W-29) ' (U-301 A specific example for these velocities is given in Fig. 14-6. The shape of the u = 0 contour is from above, s/r = (3/3[1 — (h0/p)/r\ where s/r
412 Theory of Hydrodynamic Lubrication is the angle which the vector radius r from O' to a point along the u = 0 line makes with the moving plane (Fig. 14-4). When r —» oo, s/r —> 0/3 in accordance with previous conclusions. By integrating Eq. (14-26) between hi and h2, we obtain the load capacity as Table 14-1 gives the value of P for a range of values of a with h2/0B as a parameter which is seen to have a maximum at about a = 2. The load capacity is given there as a ratio of Pa*i/Pa-i where P 1 = ML 0-1 hp(2 + h/pB) Table 14-2 gives the ratio of load capacities as obtained from this analysis and from the conventional Reynolds equation as given in Chap. 3. For the normal range of values of a the discrepancy between this theory Table 14-1. Ratio of Load Capacity of Inclined to Parallel Sliders a h2/pB \ 1 1.01 1.1 1.5 2.0 2.5 3.0 4.0 0.02 1 1.17 2.45 5.38 6.24 6.13 5.73 4.83 0.002 1 2.64 15.4 44.7 53.9 53.4 50.1 42.6 Table 14-2. Ratio of Load Capacity of Sliders Based on Eq. (14-31) to That Obtained from Conventional Theory \ h2/pB \ 1 1.001 1.01 1.1 2.0 3.0 0.02 00 60.5 6.78 1.69 1.17 1.15 0.002 00 7.00 1.57 1.07 1.02 1.02 and the conventional theory is high only at large values of h2/B, usually beyond the range of actual bearing operation. For a value of 0 around ir/2 and small values of h2/Bf solution (14-31) approaches that of classical analysis. This, however, is not true at very small convergencies of the film. In this region the classical theory tends to a zero load capacity, while here the load capacity is positive even at a = 1.
Extension of the Classical Theory 413 1 fB The center of pressure is from / XP dx given by t>Lr Jo % — 2(^o + 2a)(ln a)/(a — 1) — (3 — a)hp — 2(a -f 1) (14.32) 4(a — l)[(ln a)/(a — 1) + (a — l)h0/2a — 1] Table 14-3 gives some values of £ from this and the classical theory. Table 14-3. Centers of Pressure for ht/pB » 0.002 a x/B 1.0 1.01 1.1 1.5 2.0 2.5 3.0 4.0 New theory Conventional theory 0.333 0.500 0.438 0.501 0.496 0.510 0.537 0.566 0.569 0.587 0.605 0.608 0.630 0.633 The tangential stress on the underface of the slider is as shown in Fig. 14-7: — _ I — A 3^o\ T " k~\ ~ ~2h) and F = L [B rdx = (jh _ JB-5L) (14-33) Jo h2 \ 2a a — 1/ The coefficient of friction on the slider is / = (F + F')/W, where F' is Fig. 14-7. Friction forces in inclined slider. the tangential component of the applied load. From Eqs. (14-31) and (14-33) we thus have / = (2 ln o)/3(a - 1) + ho/2 - 1 (In a)/(a - l)2 + h„/2a - l/(a - 1) B For the runner the coefficient of friction is . _ 4/it U / ln a _ 3/?q\ Ph2 \a — 1 4a/ (14-34) (14-35)
414 Theory of Hydrodynamic Lubrication A special case of interest is the centrally loaded slider. By replacing h0 by its proper value in Eq. (14-32), we obtain the relation ht = 2[q2 + 2a - £(a2 - 1)] ln a - (a - 1)[5a + 1 -4£(a - 1)] PB (a — 1)[2f{a — 1)(a — 2) — a2 + 4a + 1] — 2[2a — £(a — 1)] ln a (14-36) The first pertinent comment about this expression is that the two terms in the numerator are nearly equal to each other, which emphasizes the strong influence of minute distortions in the shape of the fluid film. By combining Eq. (14-36) with the expressions obtained above for the case of a square slider and central loading, that is, p = ir/2 and £ = we obtain from both analytical and graphical correlations the following qualitative results: a - 1 = = Cef a = = C,P Table 14-4 gives some numerical examples, the important result again being that central loading produces a definite slider inclination. This Table 14-4. Centrally Loaded Sliders 03=ir/2) a 1.2 1.1 1.05 1.02 1.01 1.001 W/uLU 918 6,168 44,860 660,000 5,173,000 5,077,000,000 B/ht 104 367 1,369 8,197 32,310 3,188,000 f 0.11 0.06 0.03 0.01 0.006 0.0006 a 2-10"’ 3 10-* 4-10-6 2 10"6 3-10“7 3 10~10 cx 0.90659 0.91695 0.92192 0.92481 0.92575 0.92659 Ci 0.86147 0.83983 0.82070- 0.80645 0.80123 0.79633 c3 1.02586 1.04490 1.05987 1.07087 1.07490 1.07869 C< 1.07741 1.09120 1.10384 1.11356 1.11718 1.12319 Cs 1.94372 1.83394 1.77662 1.74136 1.72946 1.71868 C 6 1.89472 1.75513 1.67626 1.62612 1.60894 1.59330 C7 1.76216 1.68162 1.63793 1.61042 1.60104 1.59250 (\ 1.63224 1.47409 1.37571 1.31138 1.28913 1.26879 action is a result of the presence of pc at the leading edge. From Eq. (14-27) for £ = we have 2i = A (\ _ hi\ (vLL\ h P ap\ 2 a)\PBJB and some numerical values are given in Table 14-5. It can be seen that the actual values of pC} and thus their contribution to the load capacity,
Extension of the Classical Theory 415 are not high. Their importance lies primarily in producing convergence in the fluid film. Table 14-5. Values of pe, p0, and x0 for £ = VARIATIONS ACROSS THE FLUID FILM 14-3. Sliders with High Angle of Inclination. It was mentioned above that the solutions offered for the inclined slider approached the classical Reynolds solution for average an¬ gles of inclination. For large values of hi it will be remembered from Table 14-2 that the difference be¬ tween the new and the conventional results is appreciable. When the angle of inclination is high, the entire fluid, both ahead and astern of the bearing, has to be considered; for the motion of the fluid outside will affect the hydrodynamics of the fluid film. While this was previously done by the approximate method of assuming a constant pressure along circular arcs, a more formal solution to Eq. (14-1) will now be attempted. We shall consider the general case of an inclined slider moving with translational velocities u = C/2, v = V, and an angular velocity cj while the runner itself is moving with a velocity Ui. The coordinate axes and the configuration of the bearing are shown in Fig. 14-8. It will be recalled that = _ = Hi. U r dd V ~ dr Fig. 14-8. Slider with inclination. high angle of and thus the boundary conditions to be satisfied are d\f/ ^ 1 d\I/ rr , n
416 Theory of Hydrodynamic Lubrication A stream function satisfying Vty = 0 can be written as ^(r,0) = (A + B6 + C cos 20 + D sin 20) + r(E sin 6 + Fd cos 0 + GO sin 0) + r\H + 70 + J cos 20 -f 7C sin 20) (14-37) and the values of the constants satisfying the boundary conditions are Ci (sin a — 2a cos a) A = B = C = D = „ t/1«2 + (/20c sin a — V(a cos a -J- a) it, = 2 (sin a — a cos a) 2Ci cos a 2 (sin a — a cos a) Ci sin a 2 (sin a — a cos a) Ci cos a 2 (sin a — a cos a) in a 2 — sin2 a p _ — U1 sin2 a — Uyot sin a + V(a cos a + sin a) a2 — sin2 a q _ Ci(sin a cos a — a) + Ui{a cos a — sin a) + Va sin a a2 — sin2 a jj u) sin a bi = 4 (sin a — a COS a) — 0) cos a 2 (sin a — a cos a) -o) sin a 4 (sin a — a cos a) o) cos a 4(sin a — a cos a) / = J = A = where Ci is an arbitrary constant constituting the flux per unit length of radial distance from the origin 0. The pressure gradient along a radius is given by Ref. 6 as dp _ l du\ Tr~ r 00 \dr r r~dOj = ( — C sin 20 + D cos 20) + ^ (F cos 0 + G sin 0) — ^ 7 which upon integration yields
Extension of the Classical Theory 417 Assuming that the boundary conditions are zero at the leading and trail¬ ing edges, that is p = 0 at 0 = (a,ri) and 0 = (a,r2) and, by writing r2/ri = n, we obtain for the two constants of integration: C2 = — r-l— (F cos a + G sin a) + 4ui (in r, + ■ , In ri 1 + n y n4 - I / ~ sin a — a cos a f nri ~ 0 nVi* r . Ci = —: (F cos a + G sin a) — 2 ^ / ln n cos a [n 4- 1 n2 — 1 and thus the pressure along the slider at 0 = a is p(r) = 2„(F cos a + G sin a) I ^ - 1 + +4'‘/[A(i-¥)lnn-ln^] (i4-38) By integrating Eq. (14-38) between ri and r2, we obtain for the load capacity 2 !L=_j) -(-^[(n-D-^inn] The shearing stress as given by Ref. 6 is (dv v . 1 du)\ T', = ,l{d-r--T + -rde) = ^ (C cos 20 + Z) sin 20) — (F sin 0 — G cos 0) r r 4- 4/u(,/ cos 20 4- A sin 20) Along the plane 0 — a between ri and r2 the total drag is from F = L J" Tre dr given by P = — % (F cos a 4- G sin D a) An n —
418 Theory of Hydrodynamic Lubrication Upon substitution of the values of the constants, the equations for the load capacity and drag become p _ 2/z T U\ot sin a + U2 sin2 a — V(a + sin a cos a) 1 B [ a2 — sin2 a J cos« [(»_!)__?» J \ tt + 1/ B sin a — a cos a [ n + 1 J (14-39) F = 2yL + 2/aL Lri(a cos a — sin a) + l^Csin a cos a — a) + F sin cr — sin* a in2 a] . ln n Uia sin a + U2 sin2 a — U(a + sin « cos a)~| sin a (n — l\ a2 — sin2 a J cos a \n + 1/ + nu,Ln ^5-2 [ (n - 1) ln n1 (14-40) sin a — a cos a [ n + 1 J When the angle a is small, the pressure gradient becomes the same for both planes and the resulting load capacity becomes equal to that of classical theory, namely, ~2e-(v Ba2 \ 1 + u, -?)[- tt -f 6/ucor ~Ba* By comparing the conventional solution with the expression given in Eq. (14-39), it can be shown that the discrepancy is of the order of a2 and that even at angles as high as 30° the difference between the two solutions is of the order of only 1 per cent. While this, of course, indi¬ cates that for hydrodynamic bearings, the Reynolds assumption is quite adequate, the present results are of importance in the lubrication of disks, gears, and other machine elements that have a high angle of inclination. The approximate analysis based on the polar approximation, i.e., Eqs. (14-31) and (14-36), is adequate even for angles higher than 30°, the error again being of the order of a2. The results of the polar approximation begin to diverge from the present results when angles of the order of 7r/2 are approached. In applying Eqs. (14-39) and (14-40), it should be borne in mind that inertia forces were not taken into account and that these are not negligible for films of high convergence. From Eq. (12-4) inertia forces become of the same order as the viscous forces when Uh B p M ~ h
Extension of the Classical Theory 419 Since in our present nomenclature h — rat this can be rewritten Now in ordinary bearings a is always less than 1°. If we deal with an angle, say, of 30°, the threshold Reynolds number required to produce inertia effects is lowered to about one-thousandth of its value for con¬ ventional bearings, and any numerical example will show that, even with moderate speeds, the inertia forces soon exceed the magnitude of the viscous forces. In such cases the actual eauation to be solved is h? constant at the surfaces of journal and ^“1f bearing and (2) the normal gradient of 0 on _ v , , , ,, Fig. 14-9. Coordinate system each specified circle must equal the given for eccentric cyiinder8. velocity. The resulting pressure function must also, by physical reasoning, be single-valued. By mathematical theory any solution to the biharmonic equation can be written in either of the two following forms: these being solutions of the Laplace equation. Let us pick a coordinate system as shown in Fig. 14-9. In this diagram the origin 0 is so chosen that pU B -— « —-— where 0 = y4> i + 02 0 = (x2 + y2)<t> i + 02 V201 = V202 = 0 (14-42o) (14-426) di2 - Ri2 = d22 - R22 = s2 From this relation and from the fact that d2 — di = e (14-43) and s2 — ^2 (^2 — R\ — e)(R2 — Ri + e)(R2 + ~b e)(R2 Ri — e) (14-44)
420 Theory of Hydrodynamic Lubrication Points A and B equidistant from 0 are made the origins for the logarith¬ mic potential taken as a solution of Eqs. (14-42): <*> = C In = AP The curves on which <f> is constant are circles which include the journal and bearing. For any arbitrary circle of radius r and distance d from the origin to the center of the circle we have d2 — r2 = s2 and d«5) This form of 0, together with its derivatives, can be used to produce a large number of solutions of Eq. (14-1). Of these we shall use only those that are linear in y on either limit circle and which are single¬ valued. This yields = A ln *! + jS + y!! + B + C y) x2 + (s - y)2 x2 + (s + y)2 x2 + (s - y)2 + Dy + £(** + „* + s’) + Fy ln g-^±-g| (14-16) or in (d,y) coordinates *(d,{/) = A ln + B + c 2^-Zl) + °y + i’2 dy + Fy In (14-47) where A, B, C, D, E, and F are constants to be determined. These six constants have to be evaluated from the condition of ^ = const on either circle and also cty/dri = const. The first gives df/di/ = 0 for both circles and thus provides two boundary conditions, drp/dn provides four bound¬ ary conditions, two by setting all the coefficients containing the term l/y equal to zero and two by equating the remaining coefficients to the surface velocities of the two circles, Ricoi and R&)2. When this is done, the values of the six constants are: A = “ s2)K B = (di + s)(d2 + 8)K C = (dt - s)(d2 - s)K di ln [(d2 + s)/(d2 - s)] - d2 ln [(di + s)/(di - s)] - 2s(R22 - Rx2)/(R22 + Rx2) D = (Ri2 + Ri2) In {(di + s)(d2 - s)/(di - s)(d2 + s)] - 4se Ri2R22(Ui/Ri - U2/R2) (RiUi + R2U2) - (Ri2 + R*2)e
Extension of the Classical Theory 421 lA In [(di + s)(d2 - s)/(dl - s)(d2 + s)](RiUi + RtUt) (RS + RS) In [(di + s)(d2 - s)/(dl - s)(d2 + «)] - 4se p _ e{R\Ui -1- R2U2) (RI2 + R22) In [(di + s)(d2 - «)/(€<! - «)(d, + «)] - 4sc with 2(rfi* - di2)(flif/i + R2U2) K = (fli2 + R22)l(RI2 + /?22) In [(d! + 8)(di - s)/(d 1 - *)(d2 + «)] — 4se] , fli2fl22(E/i/fli ~ U2/R2) s(R\2 + R22){d2 — di) From the conjugate relationship between the streamlines and the pressure lines VY + i( 1/m)p = /(£ + **/) we obtain for the pressure distribution 1 _ p s(* + y) _ ^ *(« - y) _ E* 4ss M 2/2(d H- «)2 2/2(d - «) y(d2 - s2) To calculate the forces, the stress tensor of the two-dimensional field has to be first obtained. The tensors are given by o dY <Tzx = —p — 2 ii dx dy dY dx dy rm= -p + 2/» T—3- (14-49) /av av\ ^ di/2/ These can be evaluated simply from Eqs. (14-47) and (14-48) to yield 1 - i >1 s*(2rf2/ - d2 - s2) R x[2y2 + (s - 3d)y - 2sd] n 011 y2{d> - s2)2 y\d + s)3 „ x[2?/2 — (s + 3d)y + 2sd] sx(2dy - 3d’ + s2) i/2(d - s)3 j/2(rf2 - s2)2 1 - _ . . sx(2di/ - d2 — s2) , D x[2y* + (3s - d)y + 2s2] m'" y2(rf2 — s2)2 y2(d + s)3 x[2^2 - (3s + d)y + 2s2] sj-(2rf^ - d2 - s2) y2(d - s)3 ^ j/(d2 - s2")2 1 . . s[ —2dy2 + (3d2 + s2)y - 2ds2] !/2(d2 — s2)2 -2i/3 + 2(2d - s)i/2 + (3ds - 2s2 - d2)i/ + s2(d - s) + j/2(d + s)3 —2?/3 + 2(2d + s)i/2 - (3ds + 2s2 + d2)y + s2(d + s) j/2(d - s)3 , tc, s[ — 2dy2 + 4d2(/ - d(d2 + s2)] + 4 f ;f}i
422 Theory of Hydrodynamic Lubrication The equations for the components of force acting upon a circle of parameter d are wz = L<j) ^<r« p + <Tly dl ] Vu = L(f(*„;+<rnV-Zj)<ll These equations when integrated for the forces acting on the journal yield Fy = 0 Fx = W = FStcuL By replacing the constant F by its explicit value, we obtain for the load w = SryLejRiUi + R2U2) /14 (Ri2 + R22) In [(di + s)(d2 - s)/(dl - s)(d2 + s)] - 4se v where di, d2, and s are given by Eqs. (14-43) and (14-44). The torque taken as acting through the center of a circle with the parameter d is M = Lr j) |V„ - „„) dl = S*tiL(A + Fd) (14-51) For the journal the torque is then Mj = SrnL(A + Fdi) and for the bear¬ ing Mb = 87rtiL(A + Fd2)} the difference between the two being 8TnLF(d2 - di) = We The journal torque represented as a coefficient of friction from . _ M/r J W yields _ Rt^dxdi - s2) {(RS + R22) ln [(dt + s)(d2 - s)/(dl - s)(d2 - s)] - 4se] J 2Rise2(Ri2 + R22) + ft!2 + ft22 (14‘52) The above equations can be easily examined for the two extreme positions of the inner cylinder, the condition of e = 1, and the concentric case. For e = 1 we have s = 0, and since d = r and e = C, the equa¬ tion for the potential becomes *{r,y) = ^ [Rtffr - ft2) + ft2(/2(r - ft,)](r - ft,)(r - ft2) The pressure is given by 2nRitRi1 x f.„ ... /I 2 2 2 \ Ux L\~\
Extension of the Classical Theory 423 which is seen to have singularities at y = r = 0 and will yield infinite values for load capacity and torque. For the concentric case di « d2 — oo e — d\ — di — 0, and the potential is given by f, , 1 R2U2 - RiU, , . RSR^Ux/Ri + U2/R2)_ *(r) ~ 2 ft* - ft* r + ,nr The pressures here are zero. The only stress is that due to shear, which yields a torque equal to 2yRl*R2*(U1/Rl + U2/R2) M = —2irL R 22 — Ri Figure 14-10 shows the streamlines at e = 3^ for a specific case of R2/R1 = %, It is seen that a circulation is set up in the diverging portions Fio. 14-10. Streamlines between eccentric cylinders. Fig. 14-11. Isobars between eccentric cylinders. of the bearing with a direction opposite to the direction of rotation of the inner cylinder. The pressure distribution for the case of « = is shown in Fig. 14-11. First, it will be noted that here too the pressure distribu¬ tion is antisymmetrical about the line of centers, with negative pressures equal to their positive counterparts. Also, the pressure variation across the film is seen to be small even for large clearances, a thing that was a priori assumed in the Reynolds equation. Figures 14-12 and 14-13 show the load capacity and coefficient of friction calculated by these and the conventional formulas by using in the Sommerfeld equation an aintermediate’; radius of 7.
424 Theory of Hydrodynamic Lubrication When the above equations are applied to bearings, further simplifica¬ tions in the basic results of Eqs. (14-48) to (14-52) are possible. Since in bearings Ri « Ri} we have from (14-43) and (14-44) dx — = - s = - (C2 - e2)« = - (1 - «2)W e € e c and the pressure distribution becomes for t/2 = 0 6neR (7(2 — € cos 0) sin 0 V = c\2 + «2)(i - € cos ey where 0 is measured from hmin. This is the same as the standard Sommer¬ feld solution. Likewise, the load capacity and friction coefficient as 80 60 5l40 20 1 1 II 7 / / / // f / / , V / e/C 0 Q25 0.50 0.75 1.0 € Fig. 14-12. Load capacity of journal Fig. 14-13. Friction coefficient for bearings. Solutions from Eq. journal bearings. Solutions (14-50) for Rt/Ri = %. Som- from present analysis for Rt/Ri = %; merfeld solution for R = 7. Sommerfeld solution for R = 7. given by Eqs. (14-50) and (14-51) reduce under these conditions to the familiar expressions W = 12 tvUL(R/C)U (2 + €2)(1 - €2)* 1 (C\ 1 + 2cs 3 3\R) « SOURCES 1. Lewicki, W.: Theory of Hydrodynamic Lubrication in Parallel Sliding, Engineer, vol. 200, pp. 930-941, Dec. 30, 1955. 2. Lewicki, W.: Hydrodynamic Lubrication of Piston Rings and Commutator Brushes, Engineer, vol. 203, Jan. 18 and 25, 1957.
Extension of the Classical Theory 425 3. Milne, A. A.: A Contribution to the Theory of Hydrodynamic Lubrication, Wear, vol. 1, no. 1, 1957. 4. Wannier, G. H.: A Contribution to the Hydrodynamics of Lubrication, J. Appl. Math., vol. VIII, pp. 1-32, April, 1950. 5. Blok, H.: Discussion to (1), Engineer, vol. 202, p. 336, 1956. 6. Lamb, H.: “Hydrodynamics,” 6th ed., p. 579, Cambridge University Press, Cambridge, 1932.
CHAPTER 15 EXPERIMENTAL EVIDENCE Despite the abundance of test data, significant experiments on bearings are rare. The reasons for this lie both in the quality of the available experimental results and in the nature of the subject. The most common technical fault is that of not keeping the bearing variables isolated during testing. It is therefore difficult to obtain a correlation between perform¬ ance and a given parameter or to arrive at a basis for making comparisons and deducing trends. While this is oftentimes due to the incorrect and careless ways of the experimenter, much of it is a consequence of the nature of bearing operation. Each test bearing and each experiment involves many variables which must be carefully controlled. The clear¬ ances in journal and the tapers in thrust bearings are very minute dimen¬ sions, and the problem of machining, maintaining, and duplicating such dimensions at times becomes an art. The difficulties in obtaining perfect alignment introduce into all tests a degree of bearing misalignment, and residual 'unbalance tends to degenerate all steady-state tests into some form of dynamic loading. Some additional inevitable complications in bearing tests are heat transfer to the surroundings, thermal and elastic distortion of bearing and machine parts, and variations in ambient conditions. These complications could perhaps be minimized or accounted for; the factor which makes experiments on bearings particularly difficult is that of lubricant viscosity. The complications introduced by this parameter into the theory of lubrication have been discussed in preceding pages. Its ramifications in the experimental field are no less troublesome. It is almost impossible to vary a parameter during testing without simul¬ taneously varying the viscosity field of the lubricant. The only possible escape from this difficulty would be to use a fluid whose viscosity is not affected by temperature, pressure, or rate of shear. Such a lubri¬ cant, however, does not exist. It is, therefore, not surprising that reliable experiments on bearing behavior are both difficult and scarce. Even if such material were available, it would be impossible, nor would it be our intention, to present it in a single chapter such as this. What we propose is to offer some 426
Experimental Evidence 427 laboratory evidence which tends to confirm in a qualitative way some of the essential facets of hydrodynamic theory. The dynamics of the fluid film and the resultant motion of shaft center will be given prime consideration. We shall deal specifically with the shape of pressure profiles using a liquid, a gas, and a grease as lubricants; with the extent and striation of the fluid film in steady and dynamic loading; with the locus of shaft center; and with the breakdown of the laminar film into transient and turbulent flow. PRESSURE PROFILES The pressure distribution in journal bearings is qualitatively the same regardless of the operating conditions. However, the extent of the 135 105 75 45 15 0 15 45 75 105 135 Angle from lood line Fig. 15-1. Typical pressure profile in journal bearing. (After Smith and Fuller.) pressure wave, its rate of rise and decay, and both the magnitude and position of peak values vary from case to case. These characteristics will also depend on the properties of the lubricant used. The following are experimental results relating the pressure distribution to the bearing operating conditions while using a petroleum oil, air, and a grease as lubricants. 15-1. Liquid Lubricants. There is available in the literature an abun¬ dant store of experimental data relating pressure profiles to bearing parameters and operating conditions. We shall limit ourselves here to some data taken on a 3 X 3 X 0.005 in. bearing which emphasize the behavior of the pressure profile at the trailing end of the fluid film. This is shown in Fig. 15-1. The subatmospheric loop at the end of the pressure profile is typical and in line with the comments made in Chaps. 3 and 4 on the boundary conditions prevailing in journal bearings. The results presented are consistent with theoretical predictions regarding the com¬ mencement, location of maximum value, and gradients of the pressure
428 Theory of Hydrodynamic Lubrication profile. They emphasize the importance of deleting large negative pres¬ sures from analyses of incompressible lubrication; for, aside from the slight subatmospheric loop, no negative pressures were detected in any of the experiments. 15-2. Gaseous Lubricants. The principal difference between the pres¬ sure profiles of liquid and gaseous lubricants is that, while in a liquid they are negligible, negative pressures in gases are an important part of the hydrodynamic picture. This is partly because gases can exist at lower pressures than liquids but mainly because of the low values of the positive pressures in gases, which are of the order of one or two atmospheres. The effect of changing load, speed, and clearance in an air bearing is shown in Fig. 15-2 for a bearing with DXLXC = 1 X 1 X 0.0012 in. It is noteworthy that the negative pressures in all cases are little affected by any variation in operating conditions, and the only appreci¬ able effect on the positive pressures is caused by changing load. These tests were run with a bearing made of Veridia glass, and this led to the observation of humidity effects in the air film. This took the form of a band of moisture streaks extending all around the bearing across 5 to 30 per cent of the bearing length. This moisture could develop either during compression in the converging film or by the temperature drop in the diverging region. However, the tests showed relatively isothermal conditions in the film, and it is thus more likely that compression caused saturation and condensation of the water vapor present in the air. The amount of moisture can be expected to be small, and its minuteness is indicated by the fact that there was no noticeable increase in frictional drag when condensation occurred. 15-3. Grease as a Lubricant. For many years it was doubted whether a bearing using grease as a lubricant developed a hydrodynamic film, and it was the object of the first experiments to answer this basic question. One of the first of such experiments run on a DXLXC = 1 X 1 X 0.0071 in. bearing is reported in Ref. 2. The bearing, which had only one J32-in. pressure tap, was actually 2 in. long, but the journal was 1 in. long and could be moved laterally, thus placing the pressure hole in different positions with respect to the length of the journal. Circumferential pressures were measured by rotating the bearing. The load was applied through a threaded rider which could be moved along the shaft to ensure central loading. The lubricant was fed through a street ell into the top of a deep 60° channel, any excess grease being discharged through a small vent. The pressures were read on gauges whose lines were filled with
Experimental Evidence 429 Negative 0 / / \ 1 J I I V — J / V ^5^ s' r -20 y . Positive -30 psi Negative Positive 30 psi id) If) Fio. 15-2. Pressure profiles in hydrodynamic gas bearings, (a) Effect of load: speed, 52,800 rpm; clearance, 0.0012 in. (6) effect of speed: load, 16 psi; clearance, 0.0012 in. (c) effect of clearance: load, 5.7 psi; speed, 52,800 rpm. (After Cole.)
430 Theory of Hydrodynamic Lubrication Beoring degrees 0 45 90 135 0.5Z Bearing length /. 0° i —— 1 10° 50° _ ✓T _ "N —j o t o \ 1 A 80V^ \ 9C 0.5 L 120°^" /-' V r <25^- Z.— 140 V- —„ ■ N ■ \ V 150V- 0.5 L L 0 0.5 L L 0 Beoring length L Kb) Curves of longitudinal distribution 0.5 L I ^ Looc \\<=> \A" 'o.?\ V 4 d — “oT^ 0.05 w 45 90 135 Bearing degrees 45 90 135 Beoring degrees — Oil — Grease Rotation —3 [c) Curves of overage pressure [d) Curves of integrated pressure 0 45 90 135 Bearing degrees (o) Curves of circumferential distribution Area under curves represents force (/") exerted by film /"j, =35.6 lb = 88 % of lood ^grease = ^0.3 ^ = 101% of lood 80 40 ... H ■ Loaa s* / ' /1 ^ \ 1 NOv *-«— Diameter [d) Curves of integrated pressure Fig. 15-3. Pressure profiles in grease-lubricated bearings; N = 250 rpm. {After Cohn and Oren.) oil, and the grease was separated from the oil by a diaphragm. The diaphragm was a rayon-enforced synthetic rubber 0.02 in. thick cemented into the J^-in. enlargement of the tap hole and joined by a flexible line to the pressure gauge. Before a run the bearing was filled with grease, and a manually operated grease cup assured a continuous supply of lubricant. To read the film pressures, the oil pressure in the line was
Experimental Evidence 431 raised to about 10 psi below the expected film pressure. When the grease exceeded the back pressure, the gauges read the correct film pressures. Transverse pressures were taken at 11 positions, 0.1 in. apart in general and 0.05 in. apart at the edges. The number of circumferential pressures varied according to the steepness of the pressure profiles. The properties of the grease and of an oil used for comparative purposes are given in Table 15-1. Figure 15-3 gives test results which clearly show the hydro- dynamic nature of grease lubrication. The pressure profiles are flatter and extend over a wider arc than an oil film, which can be attributed to the higher resistance of grease to side leakage. Table 15-1. Properties of Test Lubricants NLGI No. 2 Grade Ball and Roller-bearing Grease Calcium and sodium soap, per cent 15 Mineral oil, per cent 85 Penetration at 77°F (ASTM), worked 268 unworked 256 Dropping point, °F 296 Values of extracted mineral oil: Kinematic viscosity, centistokes, at 100°F 52.8 at 210°F 6.59 Specific gravity at 60/60°F 0.9094 Pour point, °F 20 300 SS V Paraffin Oil Kinematic viscosity, centistokes, at 100°F 52.7 at 210°F 6.54 Specific gravity at 60/60°F 0.8973 A visual confirmation of the existence of cores, the unsheared regions discussed in Chap. 13, is provided by Ref. 3. Visualization of flow was achieved in a slider by dispersing bronze powder in the grease, and the motion was then observed through a microscope. The traveling micro¬ scope attached to the runner permitted measurement of the vertical posi¬ tion, and by differential focusing on individual particles the movement of the particles at various proportionate depths across the film could be ob¬ served. Stops were provided to permit unit horizontal displacement of the runner, and the relative linear and angular displacement of individual par¬ ticles was measured on the eyepiece graticules. Figures 15-4 and 15-5 give the basic test results. Since velocity profiles are theoretically a function only of the L/B and hi/h2 ratios, the results given here hold for any values of viscosity or absolute velocity. From theory for h\/h2 < 2 there is no reverse flow and the velocity gradients never become zero. Thus the behavior of grease is similar to that of oil, and this is illustrated in Figs. 15-4 and 15-5. For hi/h2 > 2 the oil film shows regions of reversed flow,
432 Theory of Hydrodynamic Lubrication Fig. 15-4. Displacement profile in oil- lubricated slider bearing, (a) Inlet/out¬ let film-thickness ratio a = 2; (b) inlet/ outlet film-thickness ratio a = 4; (c) inlet/outlet film-thickness ratio a = 6. Theoretical profiles; + + 4- ex¬ perimental observations. (After Milne.) (c) Fig. 15-5. Displacement profile in grease- lubricated slider bearing, (a) Inlet/out¬ let film thickness ratio a =2; (6) inlet/ outlet film-thickness ratio a * 4; (c) inlet/outlet film-thickness ratio a = 6. Theoretical profiles; + + + ex¬ perimental observations. (After Milne.) whereas the grease in these regions remains unsheared and forms stagnant cores. THE FLUID FILM; CAVITATION The angular extent of a fluid film in a journal bearing has been in the past the subject of considerable speculation, and its precise limits are still not definitely established. Many analyses assumed, as some still do, a complete film around a journal bearing. Since this assumption violated physical reality, a convenient alternative was to assume a full
Experimental Evidence 433 film throughout, and no film beyond, the converging region. Then, based on the requirements of continuity and the fact that liquids cannot endure negative stress, a refined boundary condition was established by requiring that both pressure and pressure gradient vanish at the end of the film. Further complications and greater obscurity prevailed when dynamic loading and hydrodynamic instability were treated. We now know that the limits of a fluid film cannot be precisely established without simultaneously considering the phenomena of cavitation and striation. These two phenomena upset all attempts at arriving at a neat rational boundary condition of the fluid film, and they convert the region past the hydrodynamic film into a complex system of lubricant, vapor, and air. It is about these regions that we shall try to learn from available experiments. 15-4. The Fluid Film under Steady Loading. It is perhaps instructive to begin with a general demonstration of the mechanism of cavitation as it may occur in bearings. A 6-in. steel ring 38 in. in outside diameter was run between two perspex spherical caps having a 19.75-in. radius of curvature, as shown in Fig. 15-6. One of the caps had axial pressure taps, and by rotating the cap, pressure read¬ ings could be taken anywhere on the surface. Photographs of the fluid film were taken through the other cap. Both caps were submerged in oil. When the ring was rotated at constant speed, air bubbles were drawn into the divergent clearance from the oil bath. Initially these united into a single bubble such as shown in Fig. 15-7. As the bubble increased in size, it divided into two bubbles separated by a thin film of oil. Finally, an equilibrium was reached and no further air bubbles were seen entering the clearance space. The time taken to reach equilibrium depended on the viscosity, minimum film thickness, and speed. Figure 15-8 shows the effect of speed on the size of the cavity. The variation of the fully developed cavity with film thickness is shown in Fig. 15-9. In the fully developed regions the two outer air bubbles were generally much wider than the inner bubbles. The outer air bubbles were con¬ stantly being swept to the rear of the cavitated regions where they reunited with the central air bubbles. The central air bubbles then so divided that the number of bubbles was constant for a particular set of conditions. At any given hmin the number of air bubbles increased with speed and viscosity, and for a given p the number of air bubbles increased with a decrease of hmia. i: Level of oil Bross cop o Perspex cop Fig. 15-6. Cavitation ap¬ paratus.
434 Theory of IIydrodynamic Lubrication (a) (6) (c) Fig. 15-7. Cavitation in the diverging portions of a fluid film. hmxn = 0.0106 in., n = 1 poise, u = 66.3 em/see; (a) I = 0; (/>) t = 4 min 35 sec; (c) t = 46 min 15 sec. (After Dowson.) Fig. 15-9. Relation between extent of cavitation and film thickness; n = 0.98 poise, u = 140 cm/sec, (a) hmia = 0.0066 in.; (b) hmiU = 0.0044 in.; (c) Amin = 0.0037 in. (After Dowson.)
Experimental Evidence 435 The Reynolds equation for a bearing such as used in this experiment Tr (rA* t) + m S) = ~&liU C0S *r £ (15‘1) with the film thickness given approximately by * - (15-2) where a is the radius of curvature. The outside boundary conditions are p = pa at r = rh where pa is the ambient pressure and r\ is the radius of the spherical cap. Theoretically, the boundary conditions for the cavitated regions are a zero pressure gradient and a constant pressure which is not generally the same as pa. The pressures in the cavitated regions must reach some minimum value related to the vapor pressure of the oil. Equation (15-1) has been solved numerically for the present spherical cap. In the cavitated region a constant pressure of 0.13 psi below pa was used, a value dictated by test data. A comparison between the theoretical solution and experimental results is shown in Fig. 15-10. The lower halves give the theoretical pressure distribution and the upper halves the experimental measurements. The contour diagrams show that the theoretical predictions, based on the zero pressure gradient boundary condition, are not fulfilled. In all instances the oil film rup¬ tures at a position that is downstream from theoretical predictions. This, as we shall see later, is confirmed by direct experiments with journal bearings. The above conclusion is based both on visual observation of the rupture of the oil film and on the pressure readings. Immediately upstream of the observed rupture position, a subcavity pressure trough exists. This pressure trough, along with the presence of air bubbles, upsets the theoretical zero pressure gradient condition which has already been made uncertain by the evidence of a downstream shift of the rupture point. All this suggests that surface forces at the air-oil vapor bound¬ aries are influencing the form of cavities and the boundary conditions. If the flow of oil in the cavitated regions is restricted to thin streams separated by air bubbles, then the contours of the cavities represent free streamlines along which the pressure is constant and the pressure gra¬ dients are such that the normal flow component is zero. This immedi¬ ately introduces pressures in the oil film below the cavity pressure. Thus the inclusion of free streamline requirements, surface tension, and circu¬ lation seems essential for the complete understanding of the mechanism of cavitation. Experiments6 were run on journal bearings made of Veridia glass. The lubricant used was Velocite E, which fluoresces brightly under ultraviolet
436 Theory of Hydrodynamic Lubrication Fig. 15-10. Contours of (p//xo>) X 10-3 over a spherical cap. 1. hiaia = 0.0106 in. (a) Experimental: /xo> = 3.05 g/cm sec*; (6) Theoretical: /xo) = 3.142 g/cm sec*. 2. hm\n = 0.0106 in. (a) Experimental: /xo> = 0.3 g/cm sec*; (b) Theoretical: /xo) * 9.425 g/cm sec*. 3. Amin =0.0106 in. (a) Experimental: /xo> = 6.3 g/cm sec*; (6) Theoretical /xo) = 6.283 g/cm sec*. 4. Amin = 0.0027 in. (a) Experimental : /xo> = 4.41 g/cm sec*; (6) Theoretical /xo) = 4.41 g/cm sec*. (After Dowson.)
Experimental Evidence 437 light. Both sides of the glass bearing were photographed simultaneously by using two cameras and an electronic-flash light source. To evaluate the angular limits of the oil film on the photographs, a scale was prepared by photographing an engraved shaft fitted inside the bearing with the clearance space filled with oil. The test bearings were loaded by means of cables and weights. Bearings tested were 0.984 in. in diameter with Fig. 15-11. Cavitation as a function of the Sommerfeld number. Inlot film: (a) S = 3.00; X = 1,500 rpm, P = 10 psi, pt = 0 psi, L/D = 1, C/D = 0.0012. (b) S = 0.15; X = 1,500 rpm, P = 65 psi, pt = 3 psi, L/D = 1, C/I) = 0.002. Outlet film: (c) S = 0.10; .V = 1,500 rpm, P = 105 psi, p.= 3 psi, L/D = 1, C/D = 0.002. (d) S = 0.02; X = 500 rpm, P = 100 psi, p. = 25 psi, L/D = 1, C/D = 0.0028. (After Cole and Hughes.) widths of 1, and 1 l/i in-; clearances were 0.001 and 0.002 in. Bearings with a single hole and with axial and circumferential grooves were tested, all grooving being 0.05 in. deep and located opposite the load. The qualitative behavior of the fluid film is shown in Figs. 15-11 to 15-13. Under light load, S > 10, the fluid film is complete. At S = 0.0 the film ruptures at the end. With a single oil hole the inlet film begins to neck down even before the trailing end ruptures. At low speeds the film pattern is steady, but at high speeds the filmlets fluctuate laterally and eventually combine with the incoming oil on the inlet side
438 Theory of Hydrodynamic Lubrication of the film. An increase in inlet pressure has little effect on the extent of the trailing film but makes the inlet film more complete and may even extend the full-width film upstream of the inlet hole. This basic pattern is retained even at low inlet pressures. When the oil supply is shut off P« = H psi P. = 23 psi (a) Axial groove hearing, inlet film Inlet film Outlet film (6) Circumferential groove bearing Fig. 15-12. Cavitation in grooved bearings, (a) N = 1,000 rpm; P = 60 psi; L/D = 1; C/R = 0.002. (6) N = 4,000 rpm, P = 60 psi, p. = 23 psi, L/D = 1, C/R = 0.002. (After Cole and Hughes.) completely, the menisci at the sides of the bearing recirculate the side- leakage oil, which maintains a hydrodynamic film for some time. When the menisci break down, the oil film also breaks down. Submerging the entire bearing in oil produced no appreciable change in the shape of the fluid film, suggesting that the cavitation bubbles are filled mostly with air and vapor from the lubricant and not from outside the bearing. The relation between the extent of the film and the Sommerfeld number is given in Fig. 15-13. These results are compared with theory for a full
Experimental Evidence 439 journal bearing with oil admission at hmmx and a trailing boundary con¬ dition of p = dp/dd = 0 at 0 = 62. At high values of S the film is seen to be fairly complete, extending over 2v radians. At high eccentricities the film is seen to rupture at values somewhat higher than those predicted by theory. As the photographs indicate, the fluid film ruptures into orderly filaments separated.by air and vapor, and the point of rupture occurs downstream from the location predicted by the zero pressure gradient boundary condition. The extent of the visible film is not coincident with the region of positive pressures, which indicates that the oil film z2* 'o £ 2 o S » § 2. o | 0 0.1 1 10 0.1 1 10 0.1 1 10 Sommerfeld number Sommerfeld number Sommerfeld number (o) (6) (c) Fig. 15-13. Extent of fluid film, (a) L/D - 0.6, clearance ratio = 0.0016; (6) L/D = 1.1, clearance ratio = 0.0012; (c) L/D = 1.7, clearance ratio = 0.0013. theoretical film end; theoretical film start; o experimental points. (After Cole and Hughes.) is prolonged by a shallow subatmospheric pressure loop. This would be compatible with flow continuity provided an inflow of oil from the bearing sides occurred. This inflow may well be provided by the menisci on the edges of the bearing which retain and recirculate some of the side- leakage oil. 15-5. The Fluid Film under Dynamic Loading. The results presented so far have been for steady loading. In dealing with dynamic loads the first question that arises is this: does the film rupture as with steady loads, or is the film, owing to the rapidity of loading and unloading, unable to follow these cyclic variations and therefore complete at all times? The following photographs will show that, even under dynamic loading, the fluid film undergoes cavitation and striation. Figure 15-14 shows the film with a circumferential groove bearing subjected to a rotating load of 40 psi. The film is not complete, and the ruptured zone rotates with the load. The cavitated zone consists of the familiar oil filaments, but these are less regular than with steady loading. At Theoretical eccentricity ratio Theoretical eccentricity ratio Theoretical eccentricity ratio
440 Theory of //ydrodynaviic Lubrication N, rpm 0 800 Fig. 15-14. Cavitation under conditions of dynamic loading; P = 40 psi, 03l < «/2, where w: is the load frequency, the inlet film widens in the direction of motion as it occurs in steady loading; at a>l > w/2 the inlet film narrows in the direction of motion. When ool = co/2, the film on the loaded side, judging by the low intensity of fluorescence, is very thin and is continuous only over 10 to 20° of the bearing arc, instead of the over 180° for other cases. For a single inlet hole the pattern is essentially the same except that for much of the load cycle the fluid film is actually detached from the inlet hole.
Experimental Evidence 441 N, rpm 1600 2400 LOCUS OF SHAFT CENTER An important parameter in bearing operation is the locus of shaft center; yet eccentricity measurements are quite difficult to obtain. Many elaborate schemes have been devised to measure film thickness: mechanical devices consisting of dial gauges, micrometers, and the like; electrical circuits using inductance, capacitance, or reluctance pickups; optical instruments such as microscopes and light-mirror arrangements; and pneumatic systems (primarily air gauges). For none of these sys¬
442 Theory of Hydrodynamic Lubrication tems is there good agreement between theory and experiment. This, as previously said, is not surprising; for it takes unusual precision and skill to measure minute movements of a bulky journal with thermal and elastic distortions constantly upsetting the entire reference system of the measuring equipment. It is with awareness of these difficulties that we are to look at the following results. 15-6. Steady Loading. Figure 15-15 shows the locus of shaft center measured by means of mechanical riders on levers or bell cranks attached to dial indicators. To eliminate the effects of thermal expansion of the rods, a second set of parallel rods moved the dial cases approximately — Theoretical LUIU3 IUI llll llll ICI j ——— short beoring 0 10 ^ Fig. 15-15. Locus of shaft center in full Fig. 15-16. Locus of shaft center in bearing. {After DuBois and Ocvirk.) ' elliptical bearing. the same amount the dial stems were moved by thermal expansion. The bearing tested was circular, 1% in. in diameter, with diametral clearances of 0.00232 and 0.00264 in. SAE 10 oil was admitted through a single hole opposite the load; the average oil temperatures ranged from 114 to 160°F. The speed range was 500 to 6,000 rpm, and the loads varied from 0 to 760 psi. The locus of shaft center for an elliptical bearing 8 in. in diameter and 8 in. long having a vertical clearance of 0.012 and a horizontal clear¬ ance of 0.024 in. is shown in Fig. 15-16. These data cover a range of 500 to 8,000 rpm and loads from 50 to 1,000 psi. The lubricant used has a viscosity of 39.3 centistokes at 100°F and 5.96 centistokes at 210°F with a specific gravity of 0.876. The average temperatures ranged from 120 to 200°F. The eccentricity gauges were of the mutual inductance type. The essential agreement between the theoretical and experimental loci is evident from both Fig. 15-15 and Fig. 15-16.
Experimental Evidence 443 16-7. Dynamic Loading. Most of the dynamic loading experiments described in this section are taken from Ref. 9. Two of the experi¬ ments, however, are taken from other sources; for these, testing con¬ ditions are specified in the text. Otherwise, whenever results are given without any qualification, they were obtained with the equipment and test bearing described in the following two paragraphs. 15-17. Journal locus when started at zero load; n = 1.52 centipoises. (a) Feed pressure, 20 psig; speed, 1,050 rpip. (b) Feed pressure, 3 psig; speed, 1,050 rpm. (c) Feed pressure, 3 psig; speed, 312 rpm. (After Simons.) The eccentricity gauges used for measuring the locus of shaft center under dynamic loading were of the capacitance type. These gauges measure changes in capacitance as determined by the spacing between shaft and bearing and are made part of a high-frequency oscillator circuit. The variations in capacitance cause sufficient changes in oscil¬ lator frequency to be measured by techniques of the type used in fre- quency-modulation broadcasting. The test bearing was circular, 4 in. in diameter, and 2 in. long with a circumferential groove on either side and with a diametral clearance of 0.004 in. It was loaded by two springs, one for constant loading and the (a) SHAFT AT I
444 Theory of Hydrodynamic Lubrication Fig. 15-18. Journal locus when started at 50-psi load. Sommerfeld number at equilibrium = 0.752. (After Simons.) CLEARANCE CIRCLE FULL-LOAD EQUILIBRIUM POSITION—■ —NO-LOAD EQUILIBRIUM POSITION (a) (b) CLEARANCE CIRCLE "FULL-LOAD EQUILIBRIUM POSITION NO-LOAD EQUILIBRIUM / POSITION ' (c) (d) Fig. 15-19. Shaft locus under various supply pressures. = 20.0 X 10-6 g sec/cm2; speed, GOO rpm; upward load, zero to 50 psi, gradually applied; Sm\n = 0.529. (a) Flood lubrication, = 0.49; (b) 5-psig feed pressure, «max = 0.45; (c) 10-psig feed pressure, (mH = 0.45; (d) 20-psig feed pressure, «,n»x = 0.39. (After Simons.)
Experimental Evidence 445 other actuated by a cam which produced sinusoidal loading. A rotating constant load could be imposed via a bell crank, one arm of which was attached to the load bearing and the other to the loading spring. Noncyclic Variations in Load and Speed. Figure 15-17 shows the path traveled by the journal center when started under zero load; Fig. 15-18 shows the path of the same journal when started under a load of 50 psi. In both cases the journal settled down to a fixed position in a matter of seconds. The time required to reach equilibrium varied; more rapid damping was obtained with high loads, high speeds, and low vis¬ cosity. Figure 15-18 is particularly instructive because it offers a graph¬ ically vivid verification of the locus of shaft center under steady loading. The path indicated was exactly retraced when the speed was slowly reduced and brought to zero. The lack of a spiraling movement here as opposed to the no-load conditions of Fig. 15-17 is to be noted. Figure 15-19 shows the motion of shaft center at constant speed with a slowly rising load. The shaft moves for a while along a line nor¬ mal to the load, which corresponds 0.9 © 0.7 \ 0.5 :| 0.3 C O) » 0.1 0 \y i 0 0.2 0.4 0.6 0.8 1.0 1.2 Frequency of lood opplicotion Frequency of spindle rotation 1.4 1.6 (i)Lju Fio. 15-20. Eccentricity ratio for con¬ stant rotating load. (After Simons.) Totallood, lb Fig. 15-21. Journal attitude for wl/w = 1. (After Hull.) to very light loading with perhaps a symmetrical positive and negative pressure field. When the load increases and film rupture occurs, the journal begins to move in the direction of the load. Rotating Constant Load. For all load frequencies higher than one- quarter shaft speed, or for col > co/4, the shaft, when subject to a con¬ stant rotating load, describes a circular orbit. The radii of these orbits, which also represent eccentricities, are plotted in Fig. 15-20. The peak at u)L = co/2 is another indication of bearing instability at load fre¬ quencies equal to one-half journal rotation. When col = co, the eccen-
CLEARANCE CIRCLE (rf) INOICATED CLEARANCE CIRCLE (e) INOICATED CLEARANCE CIRCLE Fig. 15-22. Shaft orbit under sinusoidal load, m = 50.0 X 10-6 g sec/cm*; speed N = 150 rpm; peak load I\ = ±78 psi; (R/C)H»K/Po) = 2.38. (a) u>P/o> = 0.25; (6) top/u> = 0.167; (c) wp/w = 0.357; (d) wp/u = 0.411, = 0.26, cmin = 0.21; (e) up/u) = 0.500, « = 0.77; (/) wp/u = 1.0, «mBX = 0.12, «miD = 0.10. (After Simons.) 446
Experimental Evidence 447 tricities are smaller than those obtained with a constant nonrotating load (o>l = 0). Another experiment was run on a 2-in.-diameter bearing 2 in. long with a circumferential groove 34 in. wide in the middle and a clearance of 0.00235 in. The load was applied by simply placing unbalanced weights on the shaft, and the displacement of the journal was measured with capacitance gauges. The significance of these results lies in provid¬ ing the phase angles between load and line of centers which are shown in Fig. 15-21. We see here that at very light loads the phase angle is about Fig. 15-23. Maximum and minimum eccentricity, for sinusoidal load. (After Simons.) 90°, but as the load increases, the phase angle is radically reduced, with the attitude angle always lagging the load vector. Sinusoidal Loads. It will be recalled from Chap. 8 that the shaft center under sinusoidal loading described orbits resembling ellipses; for wp/ a>< 34 orbit had the major axis normal to the load; for o)p/o) — 34 the orbit was circular; for wp/w > 34 the orbit had the major axis parallel to the load; and at cop/u < 34 the orbit became more complicated and departed from the quasi-elliptical shape. These results are clearly confirmed in Fig. 15-22. Most of these elliptical orbits closed in one shaft revolution, although at some high wp/w values the orbit oscillated between a larger and smaller size with a tendency to be damped out by the fluid film. Maximum and minimum eccentricities corresponding to the two semiaxes of the ellipses are shown in Fig. 15-23
448 Theory of IIydrodynamic Lubrication and again the resonant behavior at cop/co = is apparent. At suffi¬ ciently high values of cop/co the maximum eccentricities are, as in the case of a rotating load, smaller than for an equivalent constant load. This figure also indicates that at a>/>/w = the effect of load on the 1.0 0.9 0.8 0.7 W 0 1 0.6 f 05 I 0.4 U LU 0.3 0.2 0.1 0 0 0.4 0.8 1.2 1.6 20 2.4 2.8 3.2 3.6 4.0 4.4 4.8 5.2 5.4 6.0 6.4 6.8 7.2 /mo* u Fig. 15-24. Eccentricity for sinusoidal loading with wp/« = }£. (After Simons.) eccentricity is small. However, both viscosity and speed as shown in Fig. 15-24 have a definite effect on the eccentricity. Higher viscosities, and, to a greater degree, higher fre¬ quencies (of both load and shaft) have a definite damping effect. Sinusoidal Load with Super¬ imposed Constant Rotating Load. This load combination consisted of ± 18.75-psi sinusoidal load with (op/(o = % and a constant’ load rotating at shaft frequency, or (oijio = 1. The sinusoidal load was maintained at the same value while various magnitudes of the con¬ stant rotating load were super¬ imposed on it. Figure 15-25 gives the results. The noteworthy point here is that the maximum eccentricity is less than that for a sinusoidal load alone; also, the larger the imposed rotating load the thicker the oil film. 1.0 o 0.8 | 0.6 ;| 0.4 C <L> 3 0.2 0 0 5 10 15 20 25 30 35 40 45 Superimposed rotating load, psi ot frequency of shaft rotation Fig. 15-25. Maximum eccentricity for combination oscillating and rotating load. (After Simons.) Mill UJI _ Alternating lood, i 18.75 psi, applied!, at one-half frequency of shoft rotation - Shoft speed, 150 rpm — Lubricant, soe 10 oil, 85°F, 40psig — feed pressure —
Experimental Evidence 449 PjPx= 0.9 upfa Fig. 15-26. for sinusoidal loading with third harmonics. (After Shawki.) Sinusoidal Load with Superimposed Higher Harmonics. The wave¬ forms imposed in these tests were in the general form of P = Pi sin a)pt + Pn sin nupt + ^ which represents a combination of fundamental and harmonic load com¬ ponents. For this series of tests a D X L X C = 4 X 6 X 0.0064 in.
450 Theory of Hydrodynamic Lubrication = 24,000 - bearing was used. The loads could be varied to provide different phase angles and frequencies. Sample results for n — 3, and P%/P\ = 0.9, along with the load dia¬ grams, are given in Fig. 15-26. The significance of these results lies in demonstrating the existence of harmonic resonances. These resonances occur at submultiples of w/»/w = J^, that is, at wp/w = H, % f°r second-, third-, and fourth-harmonic components. With sufficiently high Pn/Pi ratios these resonances may become more critical than the fundamental resonance of (Op/it) = 3^. 15-8. Instability. One of the forms of hydrodynamic instability discussed in Chap. 7 was termed “resonant whip.” This form of vibration starts at a speed about twice the first critical speed of the system and persists from then on with a frequency of vibration slightly below the first critical fre¬ quency. This condition is illus¬ trated in Figs. 15-27 and 15-28. These data were obtained on a circular bearing with two axial grooves 90° from the load line. The dimensions of the bearing were D = 2 in., L — 2 in., C = 0.0025 in.; the loading, consisting only of journal weight, was about 8 psi; the lubricant had a viscosity of 150 Say- bolt seconds at 100°F and a viscosity index of 102, and it was used at average temperatures of from 120 to 140°F. The first critical frequency of the shaft was 6,100 rpm, and Figs. 15-27 and 15-28 show that the onset of whip was at about 11,650 rpm. This was very sharply indicated by a double-looping of the oscilloscope trace and by a rise in amplitude. The frequency of vibration was at first below the first critical frequency but eventually leveled off at a value very close to it. Whipping disappeared at two points for short intervals; at about 15,000 and 23,000 rpm. These two points correspond to a test stand reso¬ nance and to the second system critical frequency, respectively. At these two points the predominant frequency was that of journal rotation. When these two points were passed, the characteristic first critical fre¬ quency reappeared. The occurrence of half-frequency whirl in gas bearings is shown in 4,000 | 12,000 | 20,000 | 8,000 16,000 24,000 Shoft speed, rpm Fio. 15-27. Amplitude and frequency of resonant whip.
451
452 Theory of Hydrodynamic Lubrication Fig. 15-29. The effects of speed and load, as well as the presence of holes, on the onset and intensity of instability are clearly indicated by the incep¬ tion and relative magnitude of the orbits of vibration. (a) (6) (c) Fig. 15-29. Half-frequency whirl in hydrodynamic air bearings. D XL XC = 3.5 X 3.5 X 0.002 in. pa = 14.7 psi (a) No hole: P = 2 psi; last stable operation, 6,600 rpm; smallest loop, 6,720 rpm; largest loop, 7,320 rpm. (6) 0.035-in. hole on top: P = 2 psi + 60 psig external pressure in hole; last stable operation, 13,860 rpm; smallest loop, 14,250 rpm; largest loop, 15,250 rpm; (c) 0.035-in. hole on top: N = 8,000 rpm; P = 2 psi + external pressure in hole; last stable operation, 10 psig; smallest loop, 6 psig; largest loop, 0 psig. (After L. W. Winn, General Electric Company.) e 1.200 Q. U | 800 >* U c §* 400 0 1 0 Fig. 15-30. Half-frequency whirl and resonant whip. The occurrence of both half-frequency whirl and resonant whip is shown in Fig. 15-30. At speeds below twice the first system critical, the shaft is driven by a half-frequency vibration, while at speeds higher than twice the first critical the shaft is governed by the resonance of the system. 400 800 1,200 1,600 2,000 Shaft speed, rpm 2,400 2,800 TURBULENCE 15-9 Breakdown of Laminar Flow. Taylor’s analysis of fluid stability between two concentric cylinders was accompanied by a series of metic-
Experimental Evidence 453 Fig. 15-31. Appearance of Taylor vortexes. (After Taylor.) -2,200 1,000 ■2,000 -1,600 •1,200 -800 -400 u2/v 400 800 Fig. 15-32. Comparison between observed and calculated speeds at which instability first appears. (After Taylor.) 2,000 Adiabatic flow vertical annulus Width Mean radius 0.307 1,000 : 500 - . w Turbulent flow LammarX plus vortexes flow a * Laminar flow plus vortexes 2P00 1,500 ipoo 5001- Laminar XV flow Adiabatic flow, horizontal annulus Width =Q198 Mean radius Turbulent flow plus vortexes flow vortexes 100 200 300 400 500 600 700 Taylor's number low ^Laminar / VL plus vor* , , L 100 200 300 400 • 1 1 Toylor s number wfttCi/v 500 (b) (a) Wide annulus; (6) narrow annulus. (a) Fig. 15-33. Four instability regimes for air. {After Kaye and Elgar.) ulous experiments. The tests were run with 90-cm-long cylinders; the inner cylinder was made of steel, and the outer cylinder, with a radius of 4.035 cm, was made of glass. For the inner radius, values of 3.55 and 3.80 cm were used. The fluid used was deaerated water. To trace the fluid streamlines, six radial holes were drilled in the inner cylinder; through them eosin mixed with either ammonia or alcohol to approximate the density of water was injected.
454 Theory of IIydrodynamic Lubrication Transition from Laminar to Laminar plus Vortexes Re = 121 Re = 274 Laminar plus Vortexes Region Fio. 15-34. Appearance of Taylor vortexes using air. (After Kaye and Elgar.) (Continued on next page.) A qualitative confirmation of the creation of square counterrotating vortexes is shown in the photograph of Fig. 15-31. When both cylinders rotated in the same direction, with the outer cylinder faster than the inner, the motion was always stable. For values of cd2/«i < (Ri/R2)2 Fig. 15-32 gives the speeds at which vortexes began to form. To put all speeds on a common basis, they were all divided by the kinematic vis¬ cosity. The remarkable agreement between theory and experiment is evident. It is interesting to note that, if the outer cylinder is rotating in a direction opposite to the inner cylinder, the magnitude of au neces¬ sary to cause instability is greater than if the outer cylinder were at rest. If the speed at which the vortexes commence forming is maintained, the vortexes persist in a stable manner. If the speed is increased, the vigor of the circulation inside the vortexes increases without a change of shape. With further increase in speed, fully developed turbulence occurs. Another set of instructive experiments was run on a pair of cylinders using air as the fluid. Here the axial velocity of the air was an additional parameter. In these tests the outer cylinder was at rest while the inner cylinder rotated. Smoke was injected as a means of visualizing the flow, which was photographed with either a still or movie camera. Taylor vortexes formed also with air, but the experiments yielded four different
Experimental Evidence 455 Laminar Vortexes begin to form Vortexes almost formed Vortexes formed Transition from Laminar to Laminar plus Vortexes Vortexes almost formed Vortexes formed Transition from Laminar plus Vortexes to Turbulent plus Vortexes Fig. 15-34 {continued). Appearance of Taylor vortexes using air. Reynolds num¬ ber = 0. {After Kaye and Elgar.) flow regions. These regions are shown in Fig. 15-33 as a function of the Taylor number anRKC**/V and the Reynolds number 2 VC/v. In a sim¬ plified manner the Taylor number can be replaced by the speed of rotation and the Reynolds number by the axial velocity of the air. The appearance of the fluid film in these various regions is shown in the photographs of Fig. 15-34. If the cylinders run eccentrically, the onset and pattern of the Taylor vortexes seem to remain, qualitatively at least, the same. This was observed in a series of tests run on a l-in.-diameter bearing with radial
456 Theory of Hydrodynamic Lubrication clearances of from 0.0025 to 0.15 in. The length of the perspex bearing was varied from 1 to 12}^ in. and three lubricants were used: paraffin, water, and kerosene. These were dyed with fine aluminum paint for taking photographs. Tests were made with various axial flows to deter¬ mine their effect, if any, on the formation of Taylor vortexes. The results in general conformed to previous findings. Laminar flow broke down into rings of Taylor vortexes, similarly to the concentric case. In terms of the concentric clearance the effect of eccentricity is to raise the critical Reynolds number. The spacing of the vortexes was equal to the mean clearance, this being more sharply defined near the maximum X / / / / * j / i / t, ! perimc eoreti erotic t — —• — Th op ntal col loi in minor- f 50 100 150 200 250 300 Load, lb U) 100 200 300 400 500 600 Load, lb M Fig. 15-35. Eccentricity-load relation for turbulent operation, (a) Load-carrying capacity at 3,000 rpm; (6) load-carrying capacity at 7,450 rpm. (After Smith and Fuller.) clearance. As the speed was increased past the critical value, the parallel vortex rings assumed a helical and sinusoidal shape. The superposition of an axial flow was of no great significance. In the first place the axial flow itself was little affected by the presence of Taylor vortexes. Even at rotational Reynolds numbers five times the critical value, the axial flow remained the same as it was in the laminar region. High axial flow only tended to give the vortex a helical pattern. 15-10. Effect on Bearing Performance. The effect of turbulence in bearings is in Chap. 12 stated to be an increase in load capacity as well as an increase in friction. These points were investigated on a 3- by 3- by 0.0045-in. circular bearing using water as a lubricant. Eccentricity measurements on this bearing were taken with two air gauges positioned 45° with respect to the load vector, with an amplification of 2,000:1. Figure 15-35 shows the relation between eccentricity and load from
Experimental Evidence 457 2,400 rpm 3.000 rpm 3£00 rpm 4.000 rpm • 4,700 rpm x 5,600 rpm 0 6£50 rpm A 7,450 rpm Fig. 15-36. Eccentricity locus for turbulent operation. (After Smith and Fuller.) •2.0 -2.2 -2.4 5 "26 -2.8 -3.0 -3.2 2.2 2.4 2.6 2.8 3.0 3.2 3.4 3.6 3.8 In Re Fig. 15-37. Friction coefficient for laminar and turbulent operation. (After Smith and Fuller.)
458 Theory of Hydrodynamic Lubrication laminar theory and also the test points obtained here. Figure 15-36 shows the locus of shaft center, the points falling to the right of the theoretical laminar locus. The friction factor plotted in Fig. 15-37 shows the expected rise in power loss once the laminar flow breaks down, and a significant point in this plot is that the breakdown occurs at the theo¬ retical Reynolds number predicted by Taylor’s analysis. SOURCES 1. Cole, J. A.: Observations on the Performance of Air Lubricated Bearings, Conf. on Lubrication and Wear, Paper 95, London, 1957. 2. Cohn, G., and T. W. Oren: Film Pressure Distribution in Grease Lubricated Journal Bearings, Trans. ASME, vol. 71, p. 171, 1949. . 3. Milne, A. A.: On Grease Lubrication of a Slider Bearing, Conf. on Lubrication and Wear, Paper 102, London, 1957. 4. Dowson, D.: Investigation of Cavitation in Lubricating Films Supporting Small Loads, Conf. on Lubrication and Wear, Paper 9, London, 1957. 5. Cole, J. A., and C. J. Hughes: Oil Flow and Film Extent in Complete Journal ' Bearings, Proc. Inst. Mech. Engrs. {London), vol. 170, no. 17, 1956. 6. Cole, J. A., and C. J. Hughes: Visual Study of Film Extent in Dynamically Loaded Complete Journal Bearings, Conf. on Lubrication and Wear, Paper 87, London, 1957. 7. DuBois, G. B., and F. W. Ocvirk: Analytical Derivation and Experimental Evaluation of Short Bearing Approximation for Full Journal Bearings, NACA Rept. 1157, 1953. 8. Pinkus, O.: Closure to Analysis of Elliptical Bearing, Trans. ASME, vol. 78, July, 1956. 9. Simons, E. M.: Hydrodynamic Lubrication of Cyclically Loaded Bearings, Trans. ASME, vol. 74, August, 1952; R. W. Dayton and E. M. Simons: NACA Tech. Note 2544, 1951. 10. Hull, E. H.: Journal Bearing Behavior under Periodic Loading, Rept. 55-RL- 1354, General Electric Company. 11. Shawki, G. S. A.: Journal Bearing Performance for Combinations of Steady Fundamental and Harmonic Components of Load, Proc. Inst. Mech. Engrs. {London), 1958. 12. Pinkus, O.: Experimental Investigation of Resonant Whip, Trans. ASME, vol. 78, July, 1956. 13. Taylor, G. I.: Stability of a Viscous Liquid Contained between Two Rotating Cylinders, Proc. Roy. Soc. {London), A, vol. 223, pp. 289-343, February, 1923. 14. Kaye, J., and E. C. Elgar: Modes of Adiabatic and Diabatic Fluid Flow in an Annulus with an Inner Rotating Cylinder, Trans. ASME, vol. 80, April, 1958. 15. Cole, J. A.: Experiments on the Flow in Rotating Annular Clearances, Conf. on Lubrication and Wear, Paper 15, London, 1957. 16. Smith, M. I., and D. D. Fuller: Journal Bearing Operation at Super-laminar Speeds, Trans. ASME, vol. 78, April, 1956.
NAME INDEX Abramovitz, S., 67 Archibald, F. R., 135, 263 Arwas, E., 285, 327 Ausman, J. S., 176 Blok, H., 326, 425 Boeker, G. F., 285 Boyd, J., 100n., 135 Brand, R. S., 378 Brody, S., 176 Burgdorfer, A., 175, 176 Burwell, J. T., 263 Cameron, A., 67, 304, 350 Carter, G. K., 327 Charnes, A., 67, 135, 305, 402 Chou, Y. T., 378, 379 Cohn, G., 458 Cole, J. A., 458 Constantinescu, V. N., 175 Cope, W. F., 23, 304 Hahn, H. W., 263 Harrison, W. J., 176 Hays, D. I., 135, 263 Heinrich, G., 212 Hori, Jukin, 285 Hughes, C. J., 458 Hughes, W. J., 176, 212 Hull, E. H., 458 Kahlert, W., 378 Katto, Y., 176 Kaye, J., 458 Kingsbury, A., 67, 134 Ladanyi, D. J., 379 Lamb, H., 425 Lee, J. C., 67 Lewicki, W., 424 Lewis, P., 350 Licht, L., 212 Lyman, F. A., 379 Davies, R., 350 Dayton, R. W., 458 Dorr, J., 350 Dowson, D., 458 DuBois, G. B., 67, 458 Elgar, E. C., 458 Elrod, H. G., 176 Elwell, R. C., 176 Macks, E. F., 67 Maginnis, F. J., 135 Martin, H. M., 350 Meres, M. W., 135 Michell, A. G. M., 135, 305 Milne, A. A., 401, 402, 425, 458 Morgan, F., 135 Morgan, T. V., 67 Muskat, M., 135 Fedor, J. V., 134 Fend, F. A., 36 Flynn, P., 350 Fuller, D. D., 212, 458 Goldstein, S., 23 Gross, W. A., 176 Grubin, A. N., 350 459 Needs, S. J., 135 Newkirk, B. L., 285 Ocvirk, F. W., 67, 458 Oren, T. W., 458 Osterle, J. F., 176, 212, 305, 326, 327, 378, 402 Ott, H. H., 263
460 Theory of //ydrodynamic Lubrication Pai-Shih-I, 3G Paslay, P. R., 402 Piercy, N. A. W., 30 Pinkus, (>., 23, 135, 458 Poritsky, H., 36, 285, 350 Raimondi, A. A., 100n., 135, 170 Rayleigh, Lord, 67 Reid, J. C., 327 Reynolds, O., 23 Richardson, H. H., 212 Robertson, D., 285 Robinson, G. S. L., 36 Rosaus, J. J. van, 326 Saalfeld, K., 350 Saibel, E., 67, 135, 305, 326, 327, 378, 370, 402 Sassenfeld, H., 134 Sedney, R., 67 Shaw, M. C.. 67 Shawki, G. S. A., 458 Shchedrov, U. S., 350 Sheinberg, S. I., 176 Shires, G. L., 36 Simons, E. M., 458 Slezkin, N. A., 378 Slibar, A., 402 Smith, M. I., 458 Sneck, H. J., 305 Soda, N., 176 Sommerfeld, A., 67 Sternlicht, B., 23, 135, 176, 212, 285, 305, 327, 350 Swift, H. W., 263 Tao, L. N., 134, 379 Targ, S. M., 378 Taylor, G. I., 379, 458 Timoshenko, S., 327 Tower, Beauchamp, 1 Vogelpohl, G., 23 Walter, A., 134 Wannier, G. H., 425 Weber, C., 350 Weber, R. R., 212 Weilder, S. E, 23 Woinowsky-Krieger, S., 327 Wood, W. L., 67, 304
SUBJECT INDEX Acceleration effects, 365-367 Adiabatic solution, comparison with iso¬ thermal, 304 Air hammer, 207-211 Axial groove bearings, 111, 113 Bearing number, 138, 163 Bending moment, 313 Biharmonic equation, form of, 404, 407 solution to, 407, 416, 419-420 Bingham plastic, 381 Blasius friction coefficient, 24, 371 Boundary conditions, half Sommerfeld, 48, 242, 357 in journal bearings, actual, 38, 432-435 Sommerfeld, 38, 71 theoretical, 68-69 in thrust bearings, 56-57 zero pressure and pressure gradient, 46, 50, 69 Cam and follower, 332-335 Capillaries, laminar flow in, 179 turbulent compressible, 193 turbulent incompressible, 183 Cavitation, under dynamic loading, 440- 441 in spherical cap, 433-436 under steady load, 38, 437-439 width of streamlets, 84 Center of pressure, 57 Centrifugal force, 271 Circulation in journal bearing, 423 Circumferential groove bearing, 75-79, 85 Composite slider, gas, 145-149 with incompressible lubrication, 62 Compressibility number (see Bearing number) Computer methods, 79-81, 129-131, 171— 172, 299-301, 324-326 Conjugates, 421 Continuity equation, for compressible fluids, 5 for incompressible fluids, 390 with sources and sinks, 4 Convergence indicator, 81 Cores, criterion for occurrence, 431-432 experimental, 431-432 in journal bearings, 384-385 in thrust bearings, 382-384 Critical speed, of rigid rotor, 277, 280 of rotor-bearing system, 279, 280 Curved slider, 59 gas, 151 pivoted, 65-67 Damping functions, for hydrodynamic (self-acting) gas bearings, 167, 175 for hydrostatic (externally pressurized) gas bearings, 199 for incompressible lubricants, 276-277 D-C analogues, 83 Density of oils, 287 Dilatation, 2 Dynamic loading, definition of, 213 equation of, 214 experimental, 445-449 finite solutions, 257, 268-269 kinds of, 213-214 load components, 216 power consumption, 216 pressure distribution, 215-216 velocity components, 214-215, 256-266 zero pressure points of, 241 Eccentricity gages, capacitance, 443 mechanical, 442 mutual inductance, 442 Elastic body, 397 Elastic deformation, comparative results, 315 effect of, 40, 306 Electrolytic tank, 82 Elliptical bearings, 111-114 Ellipticity, 111 ratio, 111-112 Energy, internal, 288 intrinsic, 15 mechanical, 16 transported, 15-16
462 Theory of Hydrodynamic Lubrication Energy equation, approach to, 14 with conduction neglected, 22 iD dimensionless form, 22 general form of, 21 in nonintegrated form, 17-18 for thrust bearings, 204-206 Enthalpy, 287 Equation of state, 22-23 Experiments on bearings, availability of, 426 difficulties with, 426 film thickness, 441-442 flow visualization, 453, 454 with grease, 428, 431 Exponential sliders, adiabatic solution, 297-298 with incompressible lubricant, 122-124 with inertia included, 364-365 with viscoelastic lubricant, 397-401 Extent of film (see Cavitation) Falling bodies, circular plates, 226 in conical seats, 223 definition of, 213, 220 elliptical plates, 225 flat plates, 223-225 journals in bearings, 219-222 in spherical seats, 222 Feeding of lubricant, point of, 85, 92 Film thickness equations, effect on thrust bearings, 122, 124, 131, 134 in exponential sliders, 122 in fitted bearings, 52 in journal bearings, 41-42 in Michell sliders, 130, 303 in parabolic sliders, 65, 151, 307 in pivoted shoes, 319-322 ' in plane sliders, 58 in tapered land bearings, 302 Finite journal bearings, axial groove, 111- 113 circumferential groove, 75-79, 85 for dynamic loading, 257, 268-269 elliptical bearing, 111-114 partial bearings, 81-111 360° compressible, 163-175 360° incompressible, 71-79, 85-88 three-lobe bearings, 111-116 Finite thrust bearings, with deformation, 313-326 exponential, 122-124 Michell, 129-134 plane compressible, 152-154 plane incompressible, 125-129 step, 118-122 Fitted bearings, 51-53 Flexural rigidity, 310, 313 Floating-ring bearing, 53-55 Flow of fluids, compressible, constant area slot, 27 in diverging slots, 27-28 through orifices, 28-31 . general expressions, 12, 84 incompressible, parallel walls, 31-32 between concentric cylinder, 32-35 between eccentric cylinder, 35-36 Fluid film forces, for any velocity, 284 for equal radial and tangential veloc¬ ities, 269 general expression, 266 for large displacements, 281 for radial velocity only, 268 for small displacements, 275 Foil bearing, 306-308 Fourier series, 116, 119 Frequency of vibration, rotor-bearing system, 280 theoretical limit, 270 Friction, skin, 24, 26 Frictional force, general expression, 13-14 Full journal bearings, with circumferen¬ tial feeding, 75-79, 85 finite analytical solution, 71-75 finite numerical solution, 85-88 infinitely long, 42-48 infinitely short, 48-49 with large clearance, 419-424 load angle in, 87 Gas journal bearings, finite analytical solutions, 163-171 finite numerical solutions, 173-175 infinitely long, 152, 156-163, 166 Gas sliders, finite, 152-154 infinitely long, 141-145, 155 Gaussian error integrals, 191 Gears, 329-332 Geometric wedge, 290-292 Grease, properties, 431 yield value, 381, 390 Grease lubrication effect, on hydrody¬ namic bearings, 384-385 on hydrostatic bearings, 389, 394-395 on Reynolds equation, 380 on squeeze films, 387 Half-frequency whirl, definition of, 265 experiments, 450, 452 (See also Instability) Half Sommerfeld solutions, 48-49, 242, 357-358
Subject Index 463 Harmonic resonances, 449-450 Heat transfer, 39, 137-138 Hertz pressure distribution, 338, 340 Heun’s equation, 73-74 Humidity effect in gas bearings, 428 Hydrodynamic lubrication, definition of, 37 differential equation of, 9 history of, 1 ranges of, 351 Hydrostatic lubrication, advantages of, 177 assumptions of, 178 definition of, 177 effect of rotation on, 178, 187-190 principles of, 177-178 stability of, 199-200, 210-211 Hypergeometric function, 372 Inertia, averaged, 152-157, 360-364 effect on bearings, 351, 353 terms in Navier-Stokes equations, 353 Inflow wedge, equivalence of, 404 importance of, 403 streamlines in, 409 Instability, causes of, 264 equation of, 266 experiments, 450-452 frequency of, 265, 270 kinds of, 265 locus of, 284-285 mechanism of, 264 threshold of, 269 Internal energy, 288 Invariant, 390 Iteration method, 354 Katto and Soda solution, 159-163 Kingsbury bearings (see Pivoted-shoe bearings) Laplace equation, operator, 374 in polar coordinate, 120-121 in rectangular coordinate, 55, 118-119 solution of, 179-181 Lead and lag time, 198 Load angle, in full bearings, 87 in partial bearings, 50 Locus of end of pressure profile, 92 Locus of maximum pressures, 92 Locus of shaft center, in elliptical bear¬ ings, 442 in full bearings, 92, 442 in partial bearings, 93 on starting, 443-444 under turbulent conditions, 457 at various inlet pressures, 444 Meniscus, 438 Michell bearing, 303 Misalignment, 40 Mixing in the groove, effect of, 304 temperatures, 316 Navier-Stokes equations, in cylindrical coordinates, 5 for incompressible fluids, 352 for incompressible and one-dimensional bearings, 374 in rectangular coordinates, 4 Newtonian fluid, 7 Non-Newtonian fluids, 380, 390, 396-397 Normal stresses, in Navier-Stokes equa¬ tions, 2 in Reynolds equation, 7 Number of pads, optimum of, 134 Oil whip (see Resonant whip) Orbits of journal, circular, 242-243 elliptical, 243-247, 257-259 under sinusoidal load, 232, 253, 259- 261 under square wave load, 234 under suddenly applied load, 228 in two-cycle engine, 237 under unidirectional load, 218 Orifices, flow through single, 202 in series, 183, 194 Parallel plate bearings, adiabatic, 289, 292-294 gas, 139-141 with inflow wedge, 404-415 Partial journal bearings, centrally loaded, 88-91 eccentrically loaded, 94-111 infinitely long, 50-51 locus of shaft center, 93 Perfect gas equation, 22 Permeability in bearings, 55 Perturbation methods, 163-168, 391-394 Petroff’s equation, 46, 56 Phase angle, for arbitrary loads, 284 for constant rotating loads, 271 Pinion and gear, 332 Pinion and rack, 329-332 Pivoted-shoe bearings, effect of mixing in groove, 316 with elastic deformation, 313-326 finite, 128-129 infinitely long, 64-67 temperatures in, 317
464 Theory of Hydrodynamic Lubrication Plane sliders, with acceleration, 366-367 finite, 125-129 grease-lubricated, 381-385 with high angle of inclination, 415-419 with inertia, 355-357 infinitely long, 5S-59 with inflow wedge, 309-415 under turbulent condition, 373-378 Poisson’s equation, 35, 187-188, 223-225 Poisson’s ratio, 310 Polar approximation, 405 Porous bearings, 54-56 Power loss including striation, 84 Prandtl’s mixing length, 373 Pressure profile, end of, 92 experimental, 427, 429-430 in journal bearings, 38, 46 in thrust bearings, 39 Probability integral, 201 Pumping power requirement in hydro¬ static bearings, laminar compres¬ sible, 192, 193 laminar incompressible, 181-182, 193 turbulent compressible, 195-196 turbulent incompressible, 185-186, 193 Rayleigh’s criterion, 368-369 Reduced stresses, 390 Relaxation time, 396 Resilence, 269 Resistance coefficient, for laminar flow, 26 for turbulent flow, 24 Resonant whip, definition of, 265 experimental, 450-452 frequency of, 265 Reynolds equation, for compressible fluids, 163, 171 in cylindrical coordinates, 11, 130-134 in dimensionless form, 22, 70-80 for dynamic loading, 214, 238, 254, 255, 266 generalized form, 9 for incompressible fluids, 11 linearity of, 256, 266 in ph coordinates, 138, 168 for spherical slider, 435 for squeeze films, 262 techniques of solution, analytical, 69- 78, 116-117 computer, 79-81, 129-131, 171-172, 299-301, 324-326 d-c analogue, 83 electrolytic tank, 82 semianalytical, 124-128 underlying assumptions of, 6, 403 Reynolds number, 351 for axial flow, 455 in bearings, 351,.371 in capillaries, 182-183 expression for, 353 Rheodynamic lubrication, 380 Rheostatic bearings, 387 Rolling elements, boundary conditions, 330, 333 center of pressure, 347 comparative results, 349 with dry friction, 344, 349 effect of viscosity variation, 335 equation of, 329, 332 film thickness, 341 hydrodynamics of, 328 pressure distribution in, 338, 340 torques, 347 Rotating bearing, effect on boundary con¬ dition, 8 effect on performance, 10-11, 54 Rotating constant load, with journal ro¬ tation, 235-236 with no journal rotation, 219-220 Rotors, amplitude of vibration, 272 inertia force for, 273, 276 unbalance of, 271 Routh’s stability criteria, 210 Shear modulus, 396 Shear stresses, in Navier-Stokes equa¬ tions, 1-4 in Reynolds equations, 7 Sinusoidal loading, finite solutions, with rotation, 259-260 without rotation, 260-263 half Sommerfeld solution, 247-254 Sommerfeld solutions, with rotation, 231-233 without rotation, 219 Slip, 137 Sommerfeld boundary condition (see Boundary conditions) Sommerfeld number, 45, 85 Sommerfeld solutions, 42-45, 75, 79, 424 as Fourier series, 77 Sommerfeld substitution, 42 Specific pressure, 307 Spherical seat, 222 Spring function, in gas films, 168, 175, 198 for incompressible films, 276-277, 279 Spring-supported bearing, 308-313 Square-wave loading, with journal rota¬ tion, 233-235 without journal rotation, 219
Subject Index 465 Squeeze films, definition of, 213 equation of, 216 in gas bearings, 167 with grease lubrication, 385-387 with inertia included, 362 kinds of, 214 (See also Dynamic loading; Falling bodies) Stable oscillation, 285 Static equilibrium, 285 Step bearing, adiabatic, 294-297 gas, 148-149 hydrostatic, 200-207 incompressible, 60-62, 118-122 stability of, 207-211 Stream functions, 404, 407, 415 Stress tensor, 421 Striation (see Cavitation) Surface tension, 435 Tapered-land bearings, 129-134 Taylor number, 455 Taylor vortexes, appearance of, 370 criteria for onset, 368-379 in bearings, 370-371 experimental, 453-455 Temperature distribution, between con¬ centric cylinders, 34 in exponential sliders, 298 in hydrostatic bearings, 205-206 between parallel walls, 32 Thermal wedge, 288-292 Three-lobe bearings, 111-116 Threshold of instability, for bearing-rotor system, 278, 280 for fluid film resonance, 274 Thrust bearings, actual geometry of, 57 boundary conditions, 57, 116 effect of film shape, 31, 59, 134 Thrust bearings, elastic effects, 312, 314- 315 hydrodynamics of, 39 temperature effects, 286-304 Thixotropy, 381 Tilting-pad bearing (see Pivoted-shoe bearings) Time derivatives, 3 Trapezoidal integration, 81 Turbulence, in capillaries, 184 effect, on bearings, 40, 351, 373 in slider bearings, 378 experimental, 452-458 in orifices, 183, 186, 193-194 in series, 185 threshold of, 368-369 in bearings, 370-371 velocity components, 368 Unbalance, 40 Unstable oscillation, 285 Velocity components, 34-35 Vibration (see Dynamic loading; Insta¬ bility) Viscoelastic fluids, definition of, 396 effect on bearings, 401 equation for, 396 state of, 399 Viscosity, of air and liquids, 137 difficulties with, 286, 426 effect of, 286 expressions for, 287 mathematical definition of, 1,7 Volume viscosity, 3 Wedge action, 216 Weissenberg effect, 397 Whirl (see Half-frequency whirl)
UNIVERSITY OF MICHIGAN 9015 03525 28: