You are on page 1of 838
McGRAW-HILL SERIES IN MECHANICAL ENGINEERING JAOK P. HOLMAN, Southern Methodist University Consulting Fditor BARRON + Cryogenic Systems RcKERT - Introduction to Heat and Mass Transfer ECKERT AND DRAKE + Analysis of Heat and Mass Transfer ECKERT AND DRAKE + Heal and Mass Transfer HAM, CRANE, AND ROGERS - Mechanics of Machinery HARTENRERG AND DENAVIT - Kinematic Synthesis of Linkages HINZE - Turbulence JACORSEN AND AVRE - Engineering Vibrations JUVINALL - Engineering Considerations of Stress, Strain, and Strength KAYs - Convective Heat and Mass Transfer ticinry « Combustion Engine Processes MARTIN - Kinematics and Dynamics of Machines PHELAN « Dynamics of Machinery PHELAN - Fundamentals of Mechanical Design RAVEN - Automatic Control Engineering scarenen » Theories of Engineering Experimentation scnuicirrine « Boundary-Layer Theory SINGLE - Dynamic Analysis of Machines smiatEy Kinematic Analysis of Mechanisms SINGLEY - Mechanical Engineering Design stnaLey - Simulation of Mechanical Systems Ri STORCKI jgeration and Air Conditioning Boundary-Layer Theory Dr HERMANN SCHLICHTING Professor Emeritus at the Engineering University of Braunschweig, Germany Former Director of the Acrodynamische Versuchsanstalt Géttingen Translated by Dr. J. KESTIN Professor at Brown University in Providence, Rhode Island Seventh Edition McGRAW-HILL BOOK COMPANY New York - St. Louis - San Francisco + Auckland - Bogota - Ditsseldorf - Johannesburg - London » Madrid » Mexico - Montreal - New Delhi - Panama + Paris - Sto Paulo - Singapore » Sydney - Tokyo » Toronto BOUNDARY- LAYER THEORY Copy AML tights reserved, I he reproduced, stored ina ret cleetronic, ight © 1979, 1968, 1 ated it (he Unite . 1955 hy McGraw-Hill, In States of America, No part of t ¢ publicntion may ty form or by any an al, photocopying, recording, or others ise, without the prion written jeval system, or transmitted, 8, permission of the publisher, 34567800 KEKE 7832109 Library of Congress Cataloging in Publication Data Schlichting, Hermann, Boundary-layer theory. (McCraw Hill series in mechanic Translation of Grenzschic Bibliogeaphy: 17 p Includes indexes, 1. Boundary layer, 1, Title, 83 1979 629.132°3778-17794 ISBN 0.07-055334. engineering) Theorie. TLAT4.B ‘This book was set in An qua. The editor was Frank J, Cerra aud the production supervisor was John FF. Harte, First. published in the German language under the title “GRENZSCIICHT-THEORTE” and |. Verlag) Ginb First English Edition (Sccond Edition of the book) published in 1965 ion of the book) published in 1960 Karlsruhe Copyright L951 by (. Brann (vorm, G. Braunsche Hofbuchdrackerei Second English Edition (Fourth Bait ‘Third English Edition (Sixth Eslition of tho book) published in 1968 Contents List of Tables Foreword Author's Preface to the Seventh (English) Kdition Translator's Preface to the Seventh (English) dition From the Author's Preface to the First (German) Edi Introduction mn for a viscous tid rental laws of mali CHAPTER 1 Outline of (uid potion with friction a, Real antl perfect thuids b. Viscoxit c. Compressibility d. The Hagen: Poisenille equations of flow throngh a pipe e. Principle of similarity; the Reynolds and Mack munbers f. tamparison betwee the theory of perfect this aod experiment Referen CHART a. The boun ER UL Ontline of banndary-layer theory -y-layer concept nd vortex formation ¢. Turbulent flow in a pipe and ina honndary layer References CHAPTER HL. Derivation of the equations of motion nf m compressible viscous thiid (Navier-Stokes equations) a, Fundamental equations of motion and contimaty applied to fluid flow hb. General stress system in a deformable body c. The rate at which a fluid clement is strained in flow d. Relation hetween stress and rate of deformation cc. Stokes’s hypothesis £ Bulk viscosity and thermod ‘The Navier-Stokes ecuation: References nainic pressure 8 MIAPTER TV. General properties of the Navier-Stokes equitti a. Derivation of Reynolds's principle of similarity from the Navier-Stokes eqnations b. Frictionless flow ns “solutions” of the Navier-Stokes equations c. ‘The Navier-Stokes equations interpreted as vorticity transport equations d. The limiting case of very large viscosity (very small Reynolds mmber) c. ‘The limiting ease of very snuall viseons forces (very large Reynolds amibers) f. Mathematical illustration of the process of going to the limit R > co References xv xvii xix xxi 20 2 ory 24 28 39 a 47 47 49 52 58 80 tL oF 68 mW 0 72 2B Ww 7 80 82 vi Contents CHAPTER V. Fxnet solutions of the Navier-Stokes equations 83 a, Parallel flow 83 1. Parallel flow through a straight channel and Couette flow 84 2. The Hagen.Poiscnille theory of flow throngh a pipe 85 3. The flow between two concentric rotating cylinders 87 4. The suddenly accelerated plane wall; Stokes's first problem 90 5. Flow formation in Couette motion 91 6. Flow in a pipe, starting from rest 92 7. The flow near an oscillnting flat plate; Stokes's second problem 93 8. A gencral class of non-steady solutions 4 b. Other exact solutions 96 9. Stagnation in plane flow (Hieinenz. flow) 95 9a. ‘Two-dimensional non-steady stagnation flow 99 10. Stagnation in three-dimensional flow 100 LE. Flow near a rotating disk 102 12, Flow in convergent and divergent channels 107 13. Concluding remark 109 References 110 CHAPTER VI. Very slow motion Hz a. The differential equations for the case of very slow motion 12 b. Parallel flow past a sphere 13 ¢. The hydrodynamic theory of Iubrication 116 a. The flele-Shaw flow 123 References 125 Part B. Laminar boundary layers CHAPTER V1. Boundary-laycr equation for two-dimensional incompressible flow; boundary Inyer on a plate 127 a. Derivation of boundary-layer equations for two-dimensional flow 127 b. The separation of a boundary layer 131 c. A remark on the integration of the boundary-layer equations 133 @. Skin Triction 134 ©. ‘The boundary layer along a (Int. plate 135 £ Bonndary layer of higher order 144 References 148 CHAPTER VIL General proportics of the boundary-layer equations 160 a cteristics of a houndary layer on the Reynolds 1 150 b. “Shnilar” golutions of the boundary-layer equations 1n2, c. ‘Transformation of the boundary-layer equations into the hent-conduetion equation 157 d. ‘The mamentum and energy-integral equations for the boundary tayer 168 References 162 CTIAPTER 1X. Exact solutions of the stendy-state boundary-layer equations in two- dimcnsional motion 163 a, Flow past a wedge ! 164 b. Flow in a convergent channel 166 ¢. Flow pasta oylinder; aymmetrical case (Blasins serics) 168 d. Bomudary Inyer for the potential flow given by U(x) = Uo — ax” 173, c. Flow in the wake of flat plate at zero incidence 175 f. ‘The two-dimensional laminar jot 179 g. Parallel atacama in laminar flow 183 Contents h. Flow in the inlet length of a straight channel i. The method! of finite differences j. Boundary layer of sccond order References CHAPTER X. Approximate methods for the solntion of the’ two-dimensional, steady houndary-layer equations a. Application of the momentum equation to the flow past a flat plate at zero incidence b. The approximate method due to Th. von Karman and K. Pohthausen for two- dinensional flows c. Comparison between the approximate and exact solutions 1. Flat plate at zero incidence 2. Two-di:nensional stagnation Bow 3. Flow past a circular cylinder d. Further exnmples ¢. Laminor tlow with adverse pressure gi References lint; separation CHAPTER XI. Axially symmetrical and three-dimensional boundary layers a. Bxact solutions for axially symmetrical boundary layora 1. Rotation near the groimd 2. ‘The cireular jet 3. The axially symmetric wake 4: Bomdary layer on a body of revolution b. Approximate solutions for axially aymnictrie boundary layers I. Approximate aolutions for boundary Inyers on bodies which do not rotate 2. Flow in the entrance of a pipe 3. Bonndary layers on rotating bodies of revolution e. Relation between axially aynimetrical and two-dimensional boundary Inyers; Mangler's transformation . d, Three-dimensional honndary layers 1. The honndary layer on a yawed cylinder 2. Bonndary layers on other hodies References CHAPTER XL Thermal boundary layers in laminar flow a. Derivation of the energy equation b. Temperature increase through adiabatic compression; stagnation temperature ¢. Theory of similarity in heat transfer A, Exact solutions for the problem of temperature distribution in a viscous flow 1. Conette tow 2. Poisenille tlow (hrongh a channel with flat walls Boundary-Inyer simplifications: f. General properties of thermal boundary layers 1. Forced awd natnval lows Adinhatic wall 3. Analogy between heat transfer and skin friction 4. Wilect Of Prandtl number Thermal houndary layers in forced tlow 1. Parallel dow pinat @ flat plate at zero incidence Additional similar solutions of the equations for thermal boundary layers ‘Thermal boundary layers on isothermal bodies of arbitrary shape ‘Thermal boundary layers on walls with an arbitrary temperature distribution ‘Thermal boundary layers on rotationally synnnetric and rotating bodies Measurements on cylinders and other body shapes 7, ttlect of free-streait turbnlence b. Thermal boundary layers in natural flow References vii 185 187 194 198 201 201 265 265 208 271 276 277 280 282 285 285 286 286 289 292 22 300 303 309 viii Contents CHAPTER XVID. Laminar honndary layers in compressible flow a. Physical considerations b. Relation between the velacity and the temperature fields 1. Adiabatic wall 2. Hoat transfor (flat plate, dp/de = 0) c. The that plate at zcre incidence d. Bomdary layer with non-zero pressure gradient 1. Bxact. sointions: LT. The Mingworth-Stewartson trinsformation 1.2. Self-similar solutions 2. Approximate methods 6. Tuteraction between hock wave and boundary Reterencos CHAPTER XIV, Boundary-layer control in laminar flow a, Methods of boundary-layer control 1. Motion of the solid wall 2. Acecleratian of the boundary layer (blowing) 3. Suctio 4. njection of a different 5. Prevention of transi 6. Cooling of the wall 1b, Boundary-layer auction Theoretical resnits a hy the provision of suitable shapes. Laminar nerofoils 1.1. Fundamental equations 1.2. Bxact solutions 1.3. Approximate solutions 2. Exqeriuental results on suction 2.1. Increase in lift 2.2. Decrease in drag © Injection of a different gas (Bi 1. Theoretical ronnlte 1, The fundamental equations 1 xalutions, Appraximate solntion 2. Experimental results References ry homdary layers) CHAPTER NV. a. Ceneral renmets on the calculation of non-steady boundary layers 1. Bonndary-tayer equations ‘The method of snecessive approximations CC. Lin's method for periodic external flows to a series when a steady stream is perturbed slightly id olutions p-steady honndary layers cr farmation after impulsive start of n pial ease problem in acvclerated motion ian of the starting process ayer flows, : dL Experimental invest 6. Periodic hand: 1. Oscillating cylinder in tnid at vest 2 ‘Lin's theary af havmonic oscillations 3. FT flaw with samt, harmonic perturbation, nye flow Chrongh a pipe compressihle homndary tay Non-xtend 327 327 330 332 B32 333 340 340 340 B44 B52 358 372 397 399 309 399 402 402 402 403 408 408 408, 40 4 413 4s 415 415 416 420 423 425 428 23, 432 434 436 439 Contents ix 1. Boundary layer behind a moving normal shock wave 439 413 4t5 2. Blat plate at zero incidence with variable free-atream velocity and surface temperature References Part C. Transition CHAPTER XVIL_ Origin of turbutence I a, Some experimental resulls on transition from tani to turbulent ow L. ‘Trunsition in pipe flow 2. ‘Trunition in Ue boundary layer on a solid body b. Principles of the theory of stability of laminar tows 1. Introductory remarks 2, Foundation of the method of small disturbances 3. ‘The Orr-Sommerfeld equation 4. ‘The cigenvahic problem 5. General proparties of the Orr-Sommerfeld equation c. Remults of the theory of atabilily as they apply to the bountury layer on a fat plate at zero incidence 1, Some older investigations into stability 2, Calculation of the curve of neutral stability 3. Results for the flat plate a, Comparison of the theory of stability with experiment 1. Older measurentents of transition 2. Verification of the theory of stability by experiment cc. Effect of oscillating frec stream on transitis f. Concluding remark Referenees CHAPTER XVI Origin of turbntence IT 8. lft of premure gradient on transition in howndary fayer along rusnoth wale 1b. Determination of the position of the point of instability for prescribed body shape ¢. Effect of suction on transition ita boundary layer 506 4. Fleet of body forces on transition lo 1. Boundary layer on convex walle (centrifugal forces) 510 2. ‘The flow of non-homogenons fluids (stratification) plz c. Effects due to heat transfer and compressibility 14 1. Introductory remark pit 2. The ffect of heat transfer in incompressible flow Id 3. ‘The effect af compressibility f. Stability of a houndary layer in the presence of three-dimensional disturbs 1. Flow between concentric rotating cylinders 2. Boundary layers on concave walls Stability’ of three-dimensional bow g.) The intlueuce of raughness on transition 1. Introdnetory remarke 2. Single, cylindrical roughness elements 3. Distributed roughness h, Axially aymmetrical flows References ry layers Part D. Turbulent boundary layers CHAPTER XVITL Fuimdamentals of titmtent flow an a. Introductory remarks 555 x Contents b. Mean motion and fluctuations c. Additional, “apparent?” turbulent atresscs 4d Derivation’of the stress tensor of apparent turbulent friction from the Navier- Stokes equations: 6. Some mcasuremonts on Muctuating turbulent velocities £ Engorgy distribution in turbulent stroants g. Wind-tinnel turbulenes References SHAPTER XIX, ‘Theoretical assumptions for the caleulation of turbulent flows: a, Fundamental equations b. Prandtl's mixiug-length theory ¢. Further assumptions for the turbulent shearing stress . Von Kérmén's similarity hypothesis cc. Univers distribution laws 1. Von KArman’s volocity-distribution law 2. Prandtl's velocity-distribution law {Further development of theoretical hypotheses References CHAPTER XX. Tarbulent, flow through pipes a. Experimental results for smooth pipea b. Relation between law of friction and velocity distribution c. Universal velocity-distribution Inws for very large Reynolds numbers dl. Universal resistance law for smooth pipes at very large Reynolds numbers c. Pipes of non-cirowlar cross-section £ Rough pipos and equivalent. sand roughness, Other types of roughness Flow in curved pipes and diffusers Non-steady flow Unough a pipe j. Drag reduction by the addition of polymers References . & he CHAPTER XX1. ‘Turbulent boundary layers at zero prossnre gradient; flat plate; rotating disk; roughness a. The smooth flat plate 1. Resistance formula dedneed from the !/7-th-power velocity-distribution law 2. Resistance forraula deduced from the logarithmic velocity-distribution law 3. Purther refinements Effect. of finite dinensions; houndary layers in corners Honnelnry tayers with mvetion nud blowing ‘The rotating disk 1. ‘The “frco" clink : 2. ‘The disk in a housing c. ‘The rongh plate 1. The resistance formula for a uniformly rough plate 2, Measurements on single roughness elements 3. ‘Transition from a smooth to n rough sneface issible roughness 1 CHAPTER XXII, ‘The incompressible turbulent boundaty layer with pressure gradient a. Some experimental results *b. ‘The calculation of two-dimensional turbulent boundary layers 1. General remarks 2. ‘Truckenbrodt’s integral method 3. Basic equations 578 579 583 585 586 580, 687 591 504 596 596 600 602 609 G12 O15 623 626 629 630 631 635 636 637 640 643 644 645 647 047 649 O52 652 655 057 657 605 668 668, 674 671 672 O75 Contents xi 4. Quadrature for the ealculation of plane turbulent boundary layers 077 5. Application of the method 084 6. Remarks on the behaviour of the turbulent boundary layors in the presence of a pressure gradiont 7. Turbulent boundary layers with suction and injection 8. Boundary layers on cumbered walls ¢. Turbatent bounilary layers on avrofaila; maximum tft d. ‘Throc.dimensional boundary layers 1. Boundary layers on bodies of revolution 2. Boundary layers on rotating bos 3. Convergent and divergent boundary layers References CHAPTER XXIH. Turbulent boundary layers in compressible flow a. General remarks 1. ‘Turbulent heat transfer 2. The fundamental equations for compressible flow 3. Relation between the exchange coefficients for momentun and heat b. Relation between velocity and temperatnre distribution 1. The transfer of heat from a flat plate 2. The transfer of heat from rough surfaces. 3. Temperature distribution in compressible flow: c. Influence of Mach number; laws of friction 1. ‘The flat plate at zero incidence 2. Variable pressure References CHAPTER XXIV, Free turbutent flows; jets and wakes 729 a, General remarks 729 b, Estimation of the increase in width and of the decrease in velocity 73 :. Examples TBS 1. The smoothing out of a velocity discontinuity 736 2. Free jet boundary 737 3. Two-dimensional wake behind a single body 739 4. ‘The wake behind a row of bars 744 5. The two-dimensional jet 745 6. The circular jet 747 7. ‘The two-dimensional wall jet 750 4. Diffusion of temperature in free turbulent flow 782 References 755 CHAPTER XXV. Determination of profile drag 758 a. Gencrnt remarks 758 b. ‘The experimental method due to Betz 759 ¢. The experimental method duc to Jones 761 d. Calculation of profile dray 704 ©. Losses in the flow through easeades 769 1. General remarks 769 2. Influence of Reynolds number 772 3. Bifect of Mach number 715 References 777 Bibliography 780 Index of Authors 797 Subject Index 807 Abbreviations 813 List of most commonly used symbols 815 Table Table ‘Table Table Table Table ‘Table ‘able ‘Table Table ‘Table 2.1: Thickness of houndary layer, 6, at trailing edye of Mat plate nt zero in B.l: Functions ocenrring in the sohition of plane und axint cay 10.3: Comparison of exact and approxiny List of Tables Tle Vineosity conversion factors 1.2: Density, viseantty, nnd Kinematic viscosity of water and air in terms of tem: pernture 1.3: Kinematic viscosity P parallel tuebitent flow aymnuetrient How with stagnation point. Plane case from 1. Howarth [14]; axinlly symmetrical case from N, Frocsstins [3] 5.2: Values of the finctions needed for the description of the flow of a disk rotating in a finid at vest eal ag enleutated by i, M, diated at the wall and at a tnrge distance from the wall, parrow and de bs Gregg [32] ‘The fiction /(1)) for the boundary Inger along a flat plate at zero incidence, after L. Howarth [16] 10,1: Resnits of tn calculation of the boundary Inyer fora Mat plate at zero ine cidencr based on approximate theory 10.2: Ansitinry fiutetions for the npproximate enfentation of tnotinue bomuary layers, alter Holstein and Botten [5] Lo valties of tte boundary-layer para stagnation ow tae Of tawo-dimen: ateters for the pita Ud: The functions for the velocity distribution for the ease of rotation over a stationary wall, after J.B. Nydahl [ta] ‘Table 12.1; Physical constante ‘Table Table Table Table Table Table ‘Table 12.2: Dimensiontoss cocfliciont of heat transfer, at 12.01 Coofficionts of heat. transfer on 13.1: The fnnetion F(X) for the preasnre distribution along a tat plate in uh U1: Dimensiontess boundary-Inyer thickness dr and shape fretor d4/) und dimensionless adiabatic for a fiat plate at zero incidence, from equs, (12.70) and ob, wall tonperatn (12.78) The constant A in the eqrution for the ealentation of the cocfiieieat of heat, transfer in the neighbourhood of a stagnation point, after H. B, Sqnive [131] 12.4: Nunterical values of the function 1 (A) 12.5: Values of the frnction F(z) for the eatentation of a thermal bonndary layer on a nonisothermal wall: after DB. Spalding [120] tented ver (laminar), according to refs, (93, 94, 109, 126] feat plate in uatimd convection 0), neighbourhood of a shock wave, in accor dance with equa, (13.89) and (13. after N, Curls |24] for the wid velocity profiles in the initial length on a Mat, plate al vero i uniform suction, after R, Igtiseh [40] 98 1s 130 206 22 2s 228 209 295 305 36 310 319 371 3x6 xiv ‘able 16.15 Table 17.1: ‘Table 20.1; Table 20,2: Table 20.0: Table 21.2: Table 21.3: Table 22.1: Table 22.2: Table 23.1: ‘Table 24.1: List of Tables Wavelength 434 and frequeney fr 1/1 of nentral disturbances iat terms of the Reynolds nuntber R for the boundary Inyer on a ilat, plate at. zero incidence (Blusius profile). ‘Theory after W, Tollmien [99]; numerical calculations by R. Jordinson ]47] and D.R. Houston, both for parallel flow. See Figs. 16,10 and 16,11 Dependence of critical Reynolds number of velocity profiles with suction on dintensiontess suction voimne factor £. after Ulrich [243] Ratio of mean to maxinnint velocity in pipe flow in terms of the exponent n of the velocity distribution, according to equ. (20.6) Cooflicient of resistance for smooth pipes in terms of the Reynolds minrber Resistance formula for flat plate contputed from the logarithmic velocity profile in eqns, (21.14) and (21.15) Admissibte height of protuberances in terms of tte Reynolds number Examples on the ealeulation of admissible roughness from Fig. 21.16 Summary of the quantitios which occur in the equations for the ealeutation of the dimensiontess momentum thickness, Re, of the dimensionless energy thickness, Ra, and of the shape factor; see eqns. (22.1 1a, b) Summary of numerical constants that occur in the explicit equations for the catcutation of momentum and energy thickness; acc eqns, (22.15), (22.17), and (22.19) ‘The constants a and b for tho calculation of the coefficient, of heat transfor from eqn. (23.20) and of the recovery. factor from eqn. (23.27), after H. Rei- chardt [73] and J.C. Rotta [81] Power tawa for the increase in width and for the decrease in the centre-line velocity in terms of distance x for problems of free turbulent, flow 469 600 our 642 659 O61 078 679 2 734 Foreword To the First English Edition Bowndaryslayer theory is the cornerstone of our knowledge of the flow of air and other fluids of small viscosity under circumstances of interest in many engincer- ing applications. Thus many complex problems in acrodynamiecs have been clarified by a study of the flow within the boundary layer and its effect on tle gencral flow around the body, Such preblems include the variations of minimum drag and maximum lift of airplane wings with Reynolds number, wind-tunnel turbulence, and other paraieters, Even in those eases where a complete mathematical analysis is ut. present. impracticable, the boundary-layer concept. las been extraordinarily fruitful and useful, ‘The development. of boundary-layer theory during its first fifly years is a fas- cinating illustration of the birth of a new concept, its slow growth for many years the hands of its creator and his associates, its belated acceptance by others, and the subsequent almost exponential rise in the number of contributors to its further development. ‘The first decade following the classical paper of Prandtl in 1904 brought forth fewer than 10 papers by Prandtl and his students, a rate of about one paper per year, During the past year over 100 papers were published on varions aspects of boundary-layer theory and related experiments. The name of I. Schlichting first. appears it 1930 with his doctoral thesis on the subject of wake flow. Shortly there- after Schlichting devoted major effort to the problem of the stability of laminar boundary-layer flow, My own interest in the experimental aspects of boundary-Inyer flow began in the late twenties, With the appearance of Schlichting’s papers intensive attempts were maile to find the amplified disturbances preilicted by the theory. For 10 years the experimental results uot only failed to confirm this theory but supported the idea that transition resulted from the presence of turbulence in the free air stream as described in a theory set forth by G.I. Taylor. ‘Then on a well-remembered day in August, 1940, the predicted waves were seen in the flow near a flat plate in a wind tunnel of very low turbulence. The theory of stability described in the papers of Tollmien and Sclitichting was soon confirmed quantitatively as wellas qualitatively. German periodicals available in the United States after the war referred to a series of lectures by Schlichting on boundary-layer theory which had been published in 1942, This document of 279 pages with 116 figures was not available for some time. An English translation was given limited distribution as NACA Technical Memorandum No. 1217 in 1949. These lectures were completely rewritten to include material previously classified, confidential, or sccret from Germany and other countries. xvi Foreword The result was the hook of 483 pages and 295 figures published in 1951 in the German When this book became known to research workers and cdueators in States, there was an immediate request from several quarters for an Eng- lish Uranslation, since no comparable book was available in the English language. The technical content of the present, English edition is described in the author's preface. The emphasis is on the fandamental physical ideas rather than on mathe- matical refinement, Methods of theoretical analysis are set forth along with such experimental data as are pertinent to define the regions of applicability of the theoretical resnlis or to give. physical insight. into the phenomena, Acvonantical engineers and research scientists owe a debt of gratitnde to Professor Schlichting for this timely review of the present state of boundary-layer theory. Washington D.C, December 1954 Hugh L. Dryden Author’s Preface to the Seventh (English) Edition ‘The sixth (Fnglish) edition of this boot appeared in 1968; it differed very little from the fifth (Gerinan) edition of 1965. The first: (Cerman) edition of this hook was published in 1951, In the time interval between 195] and 1968 an English edition always followed 2 German edition, All translations have been prepared by Professog Kestin in an accomplished fashion, When I decided in 1975 to write a new edition of this book I came to the con- clusion that. the preceding sequence of a German edition followed by an English edition was no longer practicable, The reason for it was the heavily it <1 cost of printing. Consequently, I suggested to the two publishing companies, G. Brann in Karlsruhe and MeGvaw-Iill in New York, to produce a new edition only in the English language. I express my thanks to both Publishers for their consent. As in the previous editions, T attempted this time also to select for inchision the most importaut contributions from atnong the abundaut. crop that appeared in the incantime in the field of boundary-layer theory, without, however, altering the basic strueture of my book. I hope that the principal thrust of the book remained intact, namely the intent to emphasize and to present theoretical considerations in a form accessible to engineers. The subdivision of the book into four parts (Fundamental laws of motion of a viscous fluid; Laminar boundary layers; Transition; Turbulent boundary layers) has heen retained. Concerning the additions T wish to mention a few. Owing to the advent of large electronic computers it became possible to tackle many problems that were considered unsolvable in the past. ‘These inehre numerical solutions of the Navier-Stokes equations for moderately large Reynolds numbers (Chap. IV). nume- rical integration of the boundary-layer equations for laminar and turbulent flows (Chap. 1X), as well as the explicit: numorical integration of the Orr-Sommerfokl equation of the theory of stability of laminar boundary layers (Chap. XVI). An- other subject newly taken into account are exact solutions of the Navier-Stokes equations for the non-steady stagnation flow (Chap. V), the theory of the laminar boundary layer of second order (Chap. VI and LX). The sections on the caleutation of two-dimensional, ineompressible, turbulent boundary layers (Chap. XXII), on the stability of laminar boundary layers with compressibility and heat-transler ofleets (See. XVITe), and on losses in cascade flows (Chap. XXV) have been completely revised. xvi Author's Preface to the Seventh (English) Edition Along with this new material, T feel that I ought to mention the topics which I specifically omitted to include, T do not disonss the effect of chemical reactions on flow processes in boundary layers as they ocenr in the presence of hypersonic flow. The same applies to houndary layers in magneto-fluid-dynamies, low-density flows and flows of non-Newtonian fluids. I still thought that Tonght to refrain from giving an exposition of the statistical theory of turbulence in this edition, as in the previous ones, because nowadays there are available other, good presentations in book form. Once again, the lists of references have been expanded considerably in many chapters. The number of illustrations increased by about 65, hut 20 old ones have been omitted; the number of pages increased by abont 70. In spite of this, I hope that the original character of this book has been retained, and that it still ean provide the render with a bird's-eye view of this important branch of the physics of fluids. As I worked on the new mannsoript I once more enjoyed the vigorous assistance that T received from several of my professional colleagues, Professor K. Gersten con- tributed sections on boundary layers of second order to the part on laminar boundary layers (Secs. VITf and 1Xj). This is a special field which he successfully worked out in recent years. Professor T, K. Fanneloep contributed the completely reformulated section on the numerical integration of the boundary-layer equations included in Sec. IXi, In the part on turbulent boundary layers, Professor BE. Truckenbrodt. provided me with a new version of the largest portion of Chapter XXII on two- dimensional and rotationally symmetric boundary layers. Dr. L. M. Mack of the California Institute of Teclinology was good enough to contribute a new section on the stability of boundary layers in supersonic flow, Sec. XVIIe. Dr. J. C. Rotta thoroughly reviewed Part D on turbulent boundary layers and made many additions to it. For the Russian literature T reecived much help from Professor Mikhailov. The translation was once again entrusted to Professor J. Kestin’s competent pen. I ex- press my sincere thanks to all those gentlemen for their valuable cooperation. I should also like to repeat. my acknowledgement of the help I reecived from several professional friends when I worked on the fifth (German) edition. Naturally, their contributions have now been retained for the seventh edition. This is the ex- tensive contribution on compressible laminar boundary layers inChapter XIII written by Dr. F.W. Riegels, Professor K. Gersten’s section on thermal boundary layers in Chapter XII and Dr. J.C. Rotta’s text on compressible turbulent boundary layers in Chapter XXIII. Texpress my thanks to Frau Gerda Wolf, Frau Hilde Kreibohm and Mrs. Leslie Giacin for the careful preparation of the clear copy of the manuscript; Frau Gerda Wolf was also very helpful for me in the library. Messrs. Rotta, Hummel and Starke were kind enough to assist with the reading of the proofs. Last, but. not least, thauks are due to Verlag Braun for their willingness to accede to my’ wishes and for the pleasing appearance of the book, Goottingen, August 1978 Hermann Schlichting Translator’s Preface to the Seventh (English) Edition The present is the fourth edition in the English language of Professor IT. Schlich- ting’s ‘“Grenzschicht-Theorie”. Once again, the new edition was prepared in close cooperation with the Author whom I visited several times in Goettingen to finalize the contents and the wording. I wish to thank Professor Schlichting for his hospitality and Messrs. McGraw-I1ill for partinI financial assistance in connexion with these trips. This time there was no German printed edition and the modifications introduced by the author were transmitted directly to me. I owe a debt of gratitude to Professor I. B. Khalifa for his help in the task of proof-reading. My wife, Alicia, prepared the authors’ and the subject indexes and competently typed them under difficnlt. circumstances. My secretary, Mrs. Gincin in Providence, and Mrs. Kreibohm in Goettingen expertly typed the manuscript; I express to them my sincere thanks for their patience. Both publishers, Messrs. G. Braun of Karlsruhe and Messrs. McGraw-Hill of New York, spared no trouble, as on past occasions, in meeting our wishes regarding the production of the book, Providence, Rhode Island, August 1978 J. Kestin From Author’s Preface to the First (German) Edition Sinee about the beginning of the current century modern research in the field of fluid dyuaimies has achieved great, successes and has heen able to provide a the- oretical clarification of observed phenomena which the science of classical hydro- dynamics of the preceding century failod to do. Essentially three branches of fluid dynamics have become particularly well developed during the last fifly years; they include boundary-layer theory, gas dynamics, and acrofoil theory. The present. book is concerned with the branch known as boundary-layer theory. This is the oldest Dranch of modern fluid dynamics; it was founded by 1. Prandtl in 1904 when he succeeded in showing how flows involving fluids of very small viseosily, in pactialar water and air, the most important ones from the point of view of applications, can be made autcnable to mathemntioal aunlysis, ‘This was wchiovedl by taking the elects of friction into account only in regions where they are cssential, namely in the thin bonudary layer which exists in the immediate neighbourhood of a soliel hody. This concept mare it possible to clarify many phenomena which occur in flows and which had previously been incomprehensible. Most important. of all, it has beconte possible to subject problems connected with the occurrence of drag lo a theoretical analysis. The science of acronautical engineering was making rapil progress and was soon able to utilize these theoretical results in practical applications. It did, furthermore, pose many problems which could be solved with the aid of the new boundary-layer theory. Acronautical engineers have long since made the concept of a houndary layer onc of everyday use and it is now unthinkable to do without it. In other fields of machine design in which problems of flow occur, in particular in the design of turhomachinery, the theory of boundary layors made inuch slower progress, but in modern times these new concepts have come to the fore in such applications as well. ‘The present hook has heen written principally for engineers. It is the outcome of a course of lectures which the Author delivered in the Winter Semester of 1941/42 for the sciontific workers of the Acronautieal Research Institute in Braunschweig. The subject. matter has been utilized after the war in many special lectures held at. the Engineering University in Braunschweig for students of mechanical engineering ant physics, Dr. If, Hahnemann prepared a set of lecture notes after the first series of lectures had been given. These were read and amplified by the Author. They were subsequently published in mimeographed form by the Office for Scientific Docu- mentation (Zentrale fiir wissenschaftliches Berichtswesen) and distributed to a limited circle of interested scicntifie workers, Several years after the war the author decided completely to re-edit. this older compilation and to publish it in the form of a book. ‘The time seemed particularly propitious because it appeared ripe for the publication of a comprehensive book, and because the results of the research work carried out during the last. ten Lo twenty: years rounded off the whole field. rom Author's Preface to the First ( an) Falition xxii ‘The book is divided into four main parts. The first part contains two intro- ductory ohapters in which the fundamentals of houndary-layer theory are expounded without the use of mathematics and then proceeds to prepare the matlematical and physical justification for the theory of laminar boundary layers, and includes the theory of therinal boundary layers. The third part is concerned with the pheno- menon of transition fron laminar to turbulent flow (origin of turbulence), and the fourth part is devoted to turbulent flows. It is now possible to take the view that the theory of laminar boundary layers is complete in its main outline. ‘The physical relations have been completely clarified; the methods of calculation have been largely worked out and have, in many cases, been simplified to such an oxtent that they should present no difficulties to engineers. In discussing turbulent flows use has been made essentially only of the semi-empirical theorics which derive from Prandtl’s mixing length. It is true that according to present views these theories possess a number of shortcomings but nothing superior has so far beon devised to take their place, nothing, that is, which is useful to the enginecr. No account of the statistical theories of turbulence has been included because they have not yet attained any practical significance for engineers. As intimated in the tille, the emphasis has been laid on the theoretical treatinent of problems. An attempt has been made to bring these considerations into a form which can be easily grasped by engincers. Only a small mimber of resnits has been quoted from among the very vohiminous experimental material. They have been chosen for their suitability to give a clear, physical insight. into the phenomena and to provide direct verification of the theory presented. Some examples have been chosen, namely those associated with turbulent flow, because they constitute the foundation of the semi-empirical theory. An attempt. was made to demonatrate that essential progress is not, made through an accumulation of extensive experimental results but rather through a small number of fundamental experiments backed by theoretical considerations. Braunschweig, October 1950 - Hermann Schlichting Introduction Towards the end of the 19th century the science of fhtid mechanics began fo develop in two directions which had practically no points in common. On the one side there was the science of theoretical hydrodynamics which was evolved from Euler's equations of motion for a frictionless, non-viscous fluid and which achieved a high degree of completeness. Since, however, the results of this so-called classical science of hydrodynamics stood in glaring contradiction to experimental results —~ in particular as regards the very important problem of pressure lossea in pipes and channels, as well as with regard to the clrag of a body which moves through a mass of fluid — it had little practical importance. For tiis reason, practical cngincers, prompted by the need to solve the important problems arising from the rapid progress in technology, developed their own highly empirical science of hydraulics. The science of hydraulics was based on a large munber of experimental data and differed greatly in its methods and in its objects from the science of theoretical hydrodynamics. At the beginning of the present century L. Prandtl distinguished himself by showing how to unify these two divergent branches of fluid dynamics. He achieved a high degree of correlation between theory and experiment and paved the way to the remarkably successful development of fluid mechanics which has taken place over the past seventy years. It had been realized even before Prandtl that the discre- pancies between the results of classical hydrodynamics and experiment. were, in very many cases, due to the fact that the theory neglected fluid friction. Moreover, the complete equations of motion for flows with friction (the Navier-Stokes equa- tions) had been known for a long time. However, owing to the great mathematical difficulties connected with the solution of these equations (with the exception of a small unmber of particular cases), the way to a theoretical treatment of viscous fluid motion was barred. Furthermore, iu the case of the two most important fluids, uainely water and air, the viscosity is very small and, consequently, the forces due to viscous friction are, generally speaking, very small compared with the remaining forces (gravity and pressure forces). For this reason it was very difficult to comprehend that the frictional forees omitted from the classical theory influenced the motion of a fluid to so large an extent. In a paper on “Fluid Motion with Very Small Friction”, read before the Mathe- matical Congress in Heidelberg in 1904, L. Prandtl showed how it was possible to analyze viscous flows precisely in cases which had great practical importance. With tL. Prandtl, Uber Fliissigkcitsbowegung bei sche kleiner Roiby Congress, Heidelberg 1904. pp. 484~-451; sce also L. Prangtf( angewandten Mechanik, Hydro. wnd Acrodynamik (Col H. Schlichting and H. Gértler, vol. (1 pp. 575—584, Spi Proc. Third Intern. Math. elte Abhandlungen zur Worl gd. by W. Tolltnion, 1 jertvPain, *SE}- 2 Introduetion the aid of theoretical considerations and several siniple experiments, he proved that the flow about a solid body can be divided into two regions: a very thin layer in the neighbourliood of the body (boundary layer) where friction plays an essential part, and tho remaining region outside this layer, where friction may be neglectod. On the basis of this hypothesis Prandtl succeeded in giving a pliysically penetrating explanation of the importance of viscous flows, ving at the same time a maxi degree of simplification of the attendant inathematical diffienltics. The theoretical considerations were even then supported by simple experiments performed in a small water timnel wl Prandtl built with his own bands. Ife thus took tlie first step towards a reunification of thoory and practice. This boundary-layer theory proved extremoly fruitful in that. it provided an effective tool for the development of fluid dynainies. Sinee the beginning of the current eentury the new theory has been deve- loped at 2 very fast rate under the additional stimulus obtained from the recently founded science of acrodynamies. In a very short time it became one of the foundation stones of modern fluid dynainies together with the other very important develop- ments — the aerofuil theory and the science of gas dynamics. Jn more recent: times a good deal of attention has been devoted to studies of the mathomatical justification of boundary-layer theory. According to these, boundary- layer theory provides us with a first approximation in the framework of a more general theory designed to ealeulate asymptotic expansions of the solutions to the complete equations of motion. The problem is reduced to « so-called singular pertur- bation which is then solved by the method of matehed asymptotic expansions. Boundary-layer theory tlus provides us with a classic example of the application of the method of singular perturbation. A general presentation of perturbation methods in fhiid meclianies was prepared by M. Van Dyke. ‘The basia of these methods can be traced to L. Prandtl's carly contributions. The bomudlary-layer theory fiuda its application in the calculation of the skin- friction drag whicli acts on a body as it is moved through a fluid: for example the drag experienced by a flat plate at, zero incidence, the drag of a slip, of an acroplane wing, aireraft nacelle, or turbine blade, Boutdary-layer flow lias tiie peculiar property tut, under ecrtain conditions the flow in the immediate neighbourhood of a solid wall becomes reversed causing the boundary layer to separate from it. This is aceom- panied by a more or less pronounced formation of eddies in the wake of the body. Thus the pressure distribution is changed and differs markedly from that in a frictionless stream, The deviation in pressure distribution from the ideal is the cause of form drag, and its calculation is thus mace possible with the aid of boundary- layer theory. Boundary-layer theory gives an answer to the very important question of what shape must. a body be given in order to avoid this detrimental separation. Separation can also oceur in the internal flow through a channel and is not. confined to external flows past solid bodies. Problems connected with the flow of fluids. through the channels formed by the blades of turbomachines (rotary compressors anid turbines) can also be treated with the aid of boundary-layer theory, Furthermore, phenomena which occur at the point of maxinum fift, of an aerofoil and which are. associated with stalling can be understood only on the basis of boundary-layer tM. Van Dyke, Perturbation methods in (uid mechanics, Academic Press, 1964. Introduction 3 theory. Finally, probloms of heat transfer between a solid body and a fluid (gna) flowing past it also belong to the class of problems in which boundary-layer pheno- mena play a decisive part. At first the boundary-layer theory was developed mainly for the case of laminar flow in an incompressible fluid, as in this case the phenomenological hy pothesis for shearing stresses alreuly existed in the form of Stokes’s knw. ‘This topic was subsequently developed in a large number of research papers atl reached sucl a stage of perfection that at present the problem of laminar How cau be considered to lave been solved in its main outline, Later the theory was extended to inchide turbulent, incompressible boundary layers which are more important from the point of view of practical applications. It is true that in the ease of turbulent flows O, Rey- niolds introduced the fundamentally important concept of apparent, or virtual turbu- lent stresses as far back as 1880, However, this concept was in itself insufficient. to make the theoretical analysis of turbulent flows possible, Great progress was achieved with the introduetion of Prandtl’s mixing-length theory (1925) which, togethor with systematic experiments, paved the way for the theoretical treatment. of turbulent flows with the aid of boundary-layer theory. Howover, a rational theory of fully developed turbulent flows is still nonexistent, and in view of the extreme com- plexity of stich flows it will remain so for a considerable time. One cannot even be certain that science will ever be successful in this task. In modern times the pheno- mena which oceur in the boundary layer of a compressible flow have become the subject of intensive investigations, the impulse having licen provided by the rapid increase in the speed of flight of modern aircraft. In addition to a velocity boundary layer such flows develop a thermal boundary layer and its existence plays an im- portant part in the process of heat transfor between the {lnid and the solid body past which it flows. At very high Mach numbers, the surface of the solid wall becomes heated to high temperature owing to the production of frictional heat (“thermal barrier”), This phenomenon presents a difficult. analytic problem whose solution is important in aircraft design and the understanding of the motion of satellites. The phenomenon of transition from laminar to turbulent flow whieh is fundamen- tal for the science of fluid dynamics was first investigated at the end of the 19th cen- tury, namely by O. Reynolds. In 1914 L, Prandtl carried out, his famous experiments with spheres and suececded in showing that the flow in the boundary layer ean also be cither laminar or turbulent and, furthermore, that. the problem of separation, and hence the problem of the caleulation of drag, is governed by this transition. Theoroti- cal investigations into the process of transition from laminar to turbulent flow are based on the acceptance of Reynolds’s hypothesis that the latter occurs as a con- sequence of an instability developed by the uninar boundary layer. Prandtl initiate his theoretical investigation of transition in the year 1921; after many vain cllorte, snecess caine in the year 1929 when W. Tollmicn compnted theoretically the critical Reynolds number for transition on a flat plate at zero incidence. However, more than ten years were to pass before Tollmicn's theory could be verified through the very careful experiments performed by II. L, Dryden and his coworkors. The stability theory is capable of taking into account the effect of a number of parameters (pressure gradient, suction, Mach number, transfer of heat) on transition, ‘This theory has fonnd many important, applications, among them in the design of acrofoils of very low drag (laminar acrofoils). 4 Introduction Modern investigations in field of fluid dynamics in general, as well as in tho field of boundary-layer resoarch, are characterized hy a very close relation botween theory and experiment. The most important steps forwards have, in most cages, been taken as a result of a small number of fundamental experiments backed hy theoretical considorations. A review of the development of boundary-layer theory which stresses the mutual cross-fertilization between theory and experiment. ia contained in an article written by A. Betzf. For about, twenty years after its incoption by L. Prandtl 1904 tho bonndary-layer theory was being developed almost exchisively in his own institute in Goettingen. One of the reasons for this state of affairs may well have been rooted in the circumstance that Prandtl’s first. publication on boundary-layer theory which appeared in 1904 was very difficult to understand. This period can be said to have ended with Prandtl’s Wilbur Wright Memorial Lecture® which was delivered in 1927 at a meeting of the Royal Aeronautical Society in London, In later years, roughly since 1930, other rescarch workers, par- ticularly those in Great, Britain and in the U.S.A., also took an active part in its development. Today, the study of boundary-layer theory has spread all over the world; together with other branches, it constitutes one of the most important pillars of fluid mechanics. The first survey of this branch of scicuce was given by W. Tollinien in 1931 in two short. articles in the “Ilandbuch der Experimentalphysik” t, Shortly after- wards (1935), Prandtl published a comprehensive presentation in “Aerodynamic Theory” edited by W. F. Durand’, During the intervening four decades the volume of rescarch into this subject. has grown cnormously§. According to a review published by II. L. Dryden in 1955, the rate of publication of papers on boundary-layer theory reached one hundred per annum at that, time. Now, some twenty years later, this rate has more than tripled. Like several other fields of research, the theory of boundary layers has reached a volume which is so enormous that an individual scientist, oven one working in this ficld, cannot be expected to master all of its specialized subdivisions. [t is, therefore, right that, the task of describing it in a modern handbook has been entrtsted to several authorst, The historical develop- ment of boundary-layer theory has recently been traced by I. Tani*. t A. Botz, Zicle, Wege und lonstrulctive Auswertung der Strémungsforschung, Zeitschr, VDI 97, (1949) 253. ° 1, Prandul, Phe goweration of vortices i Muids of small viscosity (6th Wilbur Wright Memorial Lecture, 1927), J. Roy. Acro. Soc, 31, 721-741 (1927). CY. the bibliogeaphy on p. 780. Lu. Pennultt The mechautes of viscous Nude. Aerodynamic 34 208, Ber . B. Schlichting, Sore developments of boundary-layer research th the past thirty years (The ‘Third Lanchester Memorial Lecture, 1958), J. Roy. Aero. Soc. 64, 63-80 (1960). See also: Fl. Schlichting, Recent progress in boundary-layer resenroh (The 37th Wright Brothers Memorial Lecture, 1973), ALAA Journal 12, 427 - 440 (1974). 1. Tani, History of boundary-layer research, Annual Rev. of Flnid Mechanics 9, 87— ULL (1977). theory (W. F. Dorand, ed.), Vol. 3, Part A. Fundamental laws of motion for a viscous fluid CHAPTER I Outline of fluid motion with friction a. Real and perfoct Muids Most theoretical investigations in the ficld of uid dynamics are based on the concept of a perfect, i.e. frictionless and incompressible, fluid. In the motion of sueli a perfect Muicd, two contacting layers experinnee no tangential forces (shearing stresses) but act on cach other with normal forecs (pressures) only. This is equivalent, to stating that a perfect fluid offers no internal resistance to a change in shape, The theory deseribing the motion of a perfect. tid is mathematically very far developed and supplies in many cases a satisfactory description of real motions, such as ¢. g. the motion of surface waves or the formation of liquid jets in air. On the other hand the theory of perfect fluids fails completely to account for the drag of a body. In this comnexion it leads to the statement that a body which moves uniformly through a fluid which extends to infinity experiences no drag («l'Alembert’s paradox). This unacceptable result of the theory of a perfect fluid can be traced to the fact that, the inner layers of a real fluid transmit tangential as well as normal stresses, this heing also the case near a solid wall wetted by a fluid. These tangential or friction forces ina real (lui are connected with a property which is called the viscosity of the fhiid, Because of the absence of tangential foreos, on the boundary between a perfect Huid aud a solid wall there oxists, in general, a difference in rekuive tangential velocities, i. e. there is slip. On the other hand, in real fluids the existence of intor- molecular attractions causes the fluid to adlicre to a solid wall and this gives rise to shearing stresses. ‘The existence of tangential (shearing) stresses and the condition of no slip near solid walls constitute the essential differonces between a perfect and a real fluid. Certain tlnids which are of great, practical importance, such as wator and air, have very small coefficients of viscosity. In many instances. the motion of such fluids of small viscosity agrees very well with that of a perfect flnid, because in most cases the shearing stresses are very small, For this reason the existence of viscosity is completely neglected in the theory of perfect finids, mainly because this introduces a far-reaching simplification of the equations of motion, as a result of ensive mathe: matical theory becomes possible, Tt is, however, impgffatt W&@ess the fact that. 6 T. Outline of fluid motion with friction even in fluids with very small viscosities, unlike in perfect. fluids, the condition of no slip near a solid boundary prevails. ‘This condition of uo slip introduces in many es very large discrepancies in the laws of motion of perfect and real fluids. Tn par- cular, the very large discrepancy between the valuc of drag in a real and a perfect (uid has its physical origin in the condition of uo slip near a wall. ‘This book deals with the motion of fluids of small viscosity, because of the great. practical importance of the problem. During the course of the study it will become clear how this partly consistent and partly divergent behaviour of perfect and real fluids can be explained. ca b. Viscosity ‘The nature of viscosity can best be visualized with the aid of the following ex- periment: Consider the motion of a fluid between two very long parallel plates, one ich is ab rest, the other moving with a constant velocity parallel to itself, as in Fig. 1.1. Let the distance between the plates be h, the pressure being constant Vig. 1-1. Velocity distribution in a viscoun thaid between two parallel at walls (Conotic flow) throughout the fluid. Experiment teaches that the fluid adheres to both walls, so that its velocity at the lower plate is zero, and that at Ue upper plate is equal to the velocity of the plate, U. Furthermore, the velocity distribution in the tid hetween the plates is linear, so that. the fluid velocity is proportional to the distance y from the lower plate, and we have uly) = i U, (1) In order to support the motion it is necessary to apply a tangential foree to the upper plato, the foree being in equilibrinm with the frictional forees in the fluid. It is known from experiments that this force (taken per unit area of the plate) is proportional to the velocity U of the upper plate, and inversely proportional to the distance A, The frictional force per nnit arca, denoted by (frictional slicaring stress) is, therefore, proportional to U/h, for which in general we may algo substitute dufdy. The proportionality factor between 7 and dufdy, which we shall denote by j, depends on the nature of the Maid. It is small for, “thin” fluids, such a& water or alcohol, but. large in the ease of very viscous liquids, auch as oil or glycerine, ‘Thus we have obtained the fundamental relation for fluid friction in the form 2 (1.2) b, Viscosity 7 ‘The quantity jz is a property of the fluid and depends to a great extent on its tem- perature. It, is a nicasnre of the viscosity of the fluid. ‘The Inw of friction given by enn. (1.2) is known as Nercton’s law of friction. Wan. (1,2) can be regarded as the definition of viscosity. It is, however, necessary to stresa that the example considered in Mig. 1-1 constitutes a particularly simple case of fluid inotion, A geueralizition of this simple case is contained in Stokes’s law of friction (of. Chap. 11), The dimensions of viscosity can be deduced without, difficulty from eqn. (1.2)}, The shoaring stress is measured in N/in® = Pa and the velocity gradient dujdy in see}. Hence a= te ] = Pasee, in see where the square brackets are used to denote units, The above is not the only, or even the moat widely, employed unit of viscosity. Table 1.1 lists the varions units together with their conversion factors. Kqn. (1.2) is related to Hooke’s law for an clastic solid body in which case the shearing stress is proportional to the strain ‘ at r=Gy with y= 5. . (1.3) Tlere @ denotes the modulis of shear, y the change in angle between tivo lines which were originally at right angles, and € denotes the displacement, in the direction of abscissac, Whereas in the case of an clastic solid the shearing stress is proportional to the magnitude of the strain, y, experience teaches that in the case of fhrids it is proportional to the rate of change of strain dy{dlt, If we. put a (2) = (1°) # de Yay) ~ dy te]? eu Moy we shall obtain, as before, T=} because € = ut. However, this analogy is nol, complete, because the stresses in a fluid depend on one constant, the viscosity yr, whereas those in an isotropic elastic solid depend on two. Tn all fluid motions in which frictional and inertia forces interacl it is important: to consider the ratio of the viscosity, j, to the density, e, known an the kinematic viscosity, and denoted by v: wa (1A) f Wo shail consistently use in this hook the gravitational or engincoring system of units: in accordanes with international agreement the symbols kp and thf will be used to denoto the respective units of force; the corresponding units of mass will be denoted by the abbreviations kg and tb respectively. In some tables, the uvits will be those of the SI syrtem, 8 1, Outline of fluid motion with friction Table Ll. a. Absolute viscosity 1 cosity conversion factors kp sco/m? kp hej? Pa soc kp sec/m* 1 27778 x 10-4 9-8007 kp hr/m? 3,600 1 35304 x 104 Pa ace 1-0197 x 10-t 28325 x 10-8 1 kg/m hr / 22-8325 x 10-8 71-8082 x 10-8 2-7778 x 10-4 IDE sec/ft® 4-824 13562 x 10-3 47880 x 10! (bf hr/fe? 17577 x 104 4-8824 1-7237 x 105 Ih/fe see. 15175 x 10-1 42153 x 10-8 14882 kg/m br Ibf see/ft? ThE hur/ft?® Ib/ft sec. 35310 x 10° 20482 x 10-1 56893 x 10-5 6-5898 127-1 x 108 73734 x 108 20482 x 10-1 23723 x 104 1 2-0885 x 10-8 58015 x 10-8 67197 x 10-4 0-1724 x 108 58015 x 10-* 16115 x 10-8 1-8666 x 10-4 6208 x 10¢ 1 27778 x 10-4 32174 x 10! 5-358 x 107 3,000 1 11583 x 105 31081 x 10-8 8-0336 x 10-¢ 1 b, Kinematic viscosity » | wn/hr om! /ace fet/sec ftthe m?/aoe 1 3,000 1.0764 x 10! 3-8750 x 104 m*hr 27778 x 10-4 | 29900 x 10-8 1.0764 x 10! em?/sec (Stokes) | 1 x 10-4 10-36 1:0704 x 10-3 38750 ft/sec 9-2003 x 10-? |3-3445 x 10? 1 13,600 fetyhe 25806 x 10-8 |9-2003 x 107]2-n806x10 | 27778 x 10-4 1 ‘Table 1,2. Density, viscosity, and kinematic viscosity of water and air in terns of temperature Ai Water at a pressure of 0-099 MPa. (14-696 Ibfjin?) ‘Tomperature Kinematic ‘econ Kinematic Density 9 | Vi8o5tY | Viscosity | Density Viscosity | viscosity vy x 108 io » x 10 uPasec | [ft*s0c! kg/m? | #Pasee | misc 20 > _ _ 1139 156 Ib? —10 _ t 134 16-2 I21 0 1795 » 129 16:8 13-0 10 1304 126 174 13-9 20 1010 121 17-9 148 +40 655 Liz 191 IVE 60 474 106 203 19-2 80 357 0-99 215 2b7 100 283 0-94 229 244 ©. Compressibility 9 Numerical values: Iu the case of liquids the viscosity, jz, is nearly independent. of pressure and decreases at a high rate with inereasing temperature, Tn the case of gascs, to 2 first approximation, the viscosity ean be taken to be independent of pressure bnt it increases with temperature. The kinematic viscosity, », for liquids has the same type of temperature dependence as ye, because the density, @, changes only slightly with temperature, However, it the ease of gases, for which 9 decreases cousiderably with increasing temperature, » increases rapidly with temperature. Table 1.2 contains some numerical values of g, fz and y for water and aif, Table 1.3 contains some additional useful data. Liquid -— °c Glycerine : Mercury. 2... Mercury « c. Compressibility Compressibility is a measure of the change of volume of a liquid or gas under the action of oxternal forces, In this connexion we can define a modulus of elasticity, F, of volume change, by the equation dp =—B AY (1.5) Here AV/V, denotes the relative change in volume brought about by a pressure in- crease Ap. The compressibility of liquids is very small : ,g. for water # = 280,000 Ibf} which means that a pressure incroase of 1 atm (14.7 Ibf/in?) causcs a relative change in volume of about 1/20,000, i.e, 0-005 per cent, Other liquids show similar properties so that their compressibility can be neglected in most cases, and flows of liquids can be regarded as incompressible. In the case of gases, the modulus of elasticity, B, is equal to the initial pressure Po if the changes are isothermal, as can easily be deduced from the perfeet-gas law} For air at NTP (atmospheric pressure and ice-point temperature ) # = 14-7 thfjin?, which means that air is about 20,000 times more compressible than water. Similar conditions obtain for other gases. } From the perfect gas law it oan be deduced that the change in volume, AV, caused by a change of pressure Ap. satistics the relation (py Ap) (Vo HAV) = po Vo Hence Ap % — po AV/Vax 10 I, Outline of ftuid motion with friction In order to answer the question of whether it is necessary to take into account: the compressibility of gases in problems of fluid flow it-is necessary to consider whether the changes in pressure brought about by the motion of the fluid cause large changes in volume. Instead of considering volumes it is also possible to estimate the change in density, @. Owing to the conservation of mass, we can write: (Vg AV) (Q9 +40) = Vg dy: 80 that Aglag = —AV/Vq, and eqn. (1.5) can he written as Ap = B22, (5a) 2 Consequently the How of a gas cin he considered incompressible when the relative change in density remains very small, Ag/eg <1. As known from Bernonlli’s equation p -f 4g 10? = const: (w=. velocity of flow), the change of pressure, Ap, brought about. by the flow is of the order of the dynamie head g = } @ w®, so that cau. (15a) becomes Wad, (1.6) If, therefore, Ag/eq should he small compared with anity then, as seen from eqn. (1.6), wo inst. also have q/# < 1. Ithas thus heen proved that flows of gases ean be treated as incompressible, with a good degree of approximation, if the dynamic head is sinall compared with the modulus of elasticity. ‘The same result can be expressed in a different way if the velocity of sound is introduced into the equation, According to Laplace’s equation the velocity of sound is c2 = Elgg. Mence the condition Ag/eg < 1 from eqn. (1.6) ean also he written Ag _ Qo w? (2) a @ 2 pale} <}- ‘Tho ratio of the velocity of (low, 7, to the velocity of sound, ¢, is known as the Mach number Ma, (1.7) ‘The preceding argument. leads to the cor sion that compressibility ean be neglected in the treatinent of the flow of gases if 4 M?<1 (approximately incompressiblo), (1.8) i.e. if the Mach number is small compared with unity, or, in other words, if the flow velocity is small compared ‘with the velocity of sound, In the case of air, with a velocity of sound of about ¢ = 1100 ft/sce, the change in density is Ag/og = } M? = — 0.05 for a flow velocity w = 330 ft/sec. Thjs value ean be accepted as the outside limit when a gascons flow can he considered incompressible, In what follows we shall often assume the fluid to he incompressible, which ‘will restrict. the results to small Mach nnmbers, However, on several occasions, in partionlar in Chaps. X11, XL, and XXII, onr results will be extended to include compressible fluids, d. Tho Hagen-Poiseuille equations of flow through a pipe " d, The Hagen-Poiscuille equations of flow through a pipe ‘The elementary law of friction for a simple flow with shear described in Se Tb ean be applied to the important, and more general, cage of flow through a straight. pipe of circular cross-eetion having a constant diamnetor D = 2 R. The velocity the wall is zero, because of adhesion, and reaches a inaximnm on the axis, Mig. 1.2. ‘The velocity remains conatant on cylindrical surfaces which are concentric with the axis, and the mdividnal cylindrical laminag slide ov: ch other, the veloei purely axial everywhere. A motion of this Ikind is eled laminar. Ata sulbeier large distance from the ontrance section the velocity comes independent of the coordinate along the dir bution xeross the section be= jon of flow, Fig. 12. Laminar Now throagh a pipe ta De Po The fluid moves under the influence of the prossure gradient which acts in the direction of the axis, whereas in sections which are perpendicular to it the pressure may be regardod as constant. Owing to friction individual layers aot on each other with a shearing stress which is proportional to the velocity gradient dujdy, Meneco, a fluid particle is accelerated by the pressure gradient and retarded by the frictional shearing stress, No additional forces are present, and in particular inertia forces are absent, because along every streamline the velocity remains constant. In order to establish the condition of equilibrium we consider a coaxial fluid cylinder of length 1 and radius y, Fig. 1.2, The condition of equilibrinm in the z-direction requires that the pressure force (p,— p,) m y? acting on the faces of the cylinder be equal to the shear 2x yl++t acting on the circumferential area, whence we obtain raMTme, (9) Tn accordance with the law of fri ion, eqn, (1.2), we have in the present case r= —ypdujdy because » decreases with y, 80 that eqn, (1.9) leada to dumm dy pl 2 and upon integration we find , un = PIP (0-7), The constant of integration, C, is obtained from the condition of no slip at the wall Thus r= 0 aly = R, so that © == 2/4, and finally 12 J, Outline of fluid motion with friction u(y) = PP (Rey?) (1.10) ‘The velocity is seen to be distributed parabolienlly over the radiua, Fig, 1.2, and the maximum velocity on the axis heeomos = MPs pe Um = MI ‘ apni © The volume @ flowing through a scetion per nuit: time ean be easily evaluated sinec the voltine of the paraboloid of revolution is equal to 4 x base area x height. ence ” a kt Qa FP tn = EF (i Pe)- (un Bqn. (1.11) states that the volume rate of flow is proportional to the first powor of the prossure drop per unit length (p,—p,)/1 and to the fourth power of the radius of the pipe. If the mean velocity over the cross-section i = Q/x R? is introduced, eqn. (J.11) can be rewritten as i. Py Pe = BB ay. (1.12) Fiqu. (1.11) was first deduced by G. Tagen [6] and shortly afterwards by J. Poi- souille [T1L 1t is knewn as the Hagen-Poiseuille equatien Gf laminar flow through a pipe. Haqn. (1.11) cau he utilized for the experimental determination of the viscosity, The method consists in the measurement of the rate of flow and of the pressure drop across a fixed portion of « capillary tube of known rashus. Thus cnough dala are provided to determine y from equ. (EL). The type of flow to which eqns. (1.10) and (1.11) apply exists in reality only for relatively sinall radii and flow velocities. For larger velocities and radii the character of the motion changes completely: the pressure drop ecases to be proportional to the first power of the mean velocity as indicated by equ. (1.12), but becomes approxi- mately proportional lo the second power of u. The velocity distribution across a section becomes much more uniform and the well-ordered laminar motion is replaced by a flow in which irregular and fluctuating radial and axial velocity com- ponents are superimposed on the niain motion, so that, consequently, intensive mixing in a radial direction takes place. In such cases Newton’s law of friction, eqn. (1.2), ceases to he applicable. This is the caso of turbulent flow, to he discussed in great detail later in Chap. XX. Mach numbers ¢. Principle of similarity; the Reynolds The type of {luid notion discussed in the! preceding Seet lo moved under the influence of forces only, inertia forces being every where equal to zero, Ina di channel fui particles are acted upon by inertia forces in add friction forces. n was very simple nal and pressure rgent or convergent, n to pressure and e. Principle of similarity; the Reynolds and Mach numbers 13 In the present section we shall endeavour to answer a very fundamental question, namely that concerned with the conditions under which flows of different, fluids about two gcometrically similar bodics, and with identical initial (low directions display goomctrically similar streamlines, Such motions which have geomotzically, similar streamlines are called dynamically similar, ov similar flows. For two flows about geometricully sinilar bodies (e.g. about two spheres) with different fluids, different velocitics and different to be similar, it_is evidently necessary that the following. co: ied ; at all goometrically similar points the forces acting on a fluid particle must bear a fixed ratio at every instant of time," ; We shall now consider the important case when only frictional and inertia foroes are present. Elastic forees which may be due to changes in volume will be excluded, i.e, it will be assumed that the flnid is meompressible. Cravitational forces will also be excluded so that, consequently, free surfaces are not adinittod, and in the interior of the fluid the foree of gravity is assumed to he balanced by buoyancy. Under these assumptions the condition of similarity is satisfied only if at all corresponding points the ratio of incrtia and friction forees is the same. In a motion parallel to the z-axis the inertia foree per unit volume has the magnitude of g Du/Dt, where u denotes the component of velocity in the x-direction and 1/Dt denotes the substantive derivative, In the case of steady flow we can replace it by 9 dujdx «dxflt = 9 u dufax, where dujdx denotes the change in velocity with position. Thus the inertia force per unit volume is equal to gu dujax. Vor the fri tion force it is easy to deduce an expression from Newton’s law of friction, eqn, (1.2). Considering a fluid particle for which the z-direction coincides with the direction of motion, Fig. 1.3, it is found that the resultant of shearing forces is equal to (c+ & dy) ded: —rdzdz = # dzdy de. Hence the friction force per unit volume is equal to dx/dy, or by eqn, (1.2), to #0?u/dy?. Consequently, the condition of similarity, i.e. the condition that at all corres- ponding points the ratio of the inertia to the friction fore must be constant, can be written as: Fig. 1.3. Frictional forces acting on, fluid particle It is now necessary to investigate how these forces are changed when the magnitudes which determine the flow are varied. The latter include the density 9, the viscosity }, a representative velocity, e.g. the free stream velocity V, and a characteristic linear dimension of the body, c.g. the diameter d of the sphere. 14 1. Outline of fluid motion with frietion ‘The velocity w at some point in the velocity ficld is proportional to the free stream velocity V, the velocity gradient: dujax is proportional to V/d, and similarly @u/2y? is proportional to V/d®, Hence the ratio Incrtin force _ gueufax _ po Vd _ o Vd Friction fore peujdy? ~ Vidi Therefore, the condition of similarity is satisfied if the quantity @ V dj has the same valne in both flows. The quantity 9 V d/y, which, with y/o = ¥, can also be written as Vd/r, is a dimensionless mimber because it is the ratio of the two forces. 1t is Inown as the Reynolds naamber, R. Thus two flows are similar when the Reynolds number _ eld _ va R= moo (1.13) is cqual for both. This principle was first enunciated by Osborne Reynolds{12] in connexion with his investigations into the flow through pipes and is known as Reynolds's principle of similarity. ‘The fact that the Reynolds number is dimensionless ean he at once verified rectly by considering the dimensions Ibf see? ft. itt” |. v 4 Fence @ Vd _ Ibfacc? ft em Ree which proves that the Reynolds number is, in fact, dimensionless. Method of indices: Instead of the consideration of the condition of dynamic similarity, Reynolds's principle can also be dedneed by considering dimensions by the incthod of indiees. In this connexion use is made of the observation that, all physical laws must be of a form which is independent of the particular system of inils employed. In the case under consideration the physical quantities which determine the flow ure: the free stream velocity, V, a representative linear dimen- sion of the body, d. as well ns the density, g, and the viscosity, z. We now ask whether there exists a combination of these quantities in the form Pa dl ov pw", which would be dimensionless. If F denot combination will be obtained if Ve dior po foree, L length and T time, a dimonsionless _ FOLOTo. to. Without restricting the generality of the argument. it, is perm valne of unity to one of the fear indices «, 8, y, 5, because a a dimensionless quantity is also dimensionless, Assuming. ble to assign the y arbitrary power of 1, we obtain ee bane CUED (ED peuor, 15 ¢. Principle of similarity; the Reynolds and Mach 1 Equating the exponents of L,T, and F on both sides of the expression we obtain three equations: : yr 50 L: lyf dy -26--0, Tr 4 4 yp 60 the soblition of whi ‘This shows that there exists a unique dimensionless combination of the four quanti- ties V,d,@ and jz, namely the Reynolds number R Dimensionless quantities: ‘Tho reasoning followed in the preceding derivation of the Reynolds number can be extended to include the case of different Reynolls numbers in the consideration of the velocity field and forees (normal and tangential) for flows with geometrically similar boundaries. Let the position of a point, in the space around the geometrically similar bodies be indicated by the coordinates x, y z; then the ratios a/d, y/d, 2/d are its dimensionless coordinates. ‘Tho velocity com- poncuts are inade dimensionless by referring them to the free-siream velocity V, thus 4/V, »/V, w/V, and the normal and shearing stresses, p and t, ean be made dimnen- sionless hy referring thom to the donble of the dynatnic head, i. ¢. to @ V? thus: p/g F? and t/g V®. The previously enunciated principle of dynamical similarity can he expres- sed in an alternative form by asserting that for the two geometrically similar systems. with oqual Reynolds numbers the dimensionless quantities u/V,..., ple V? and 1/o V2 depend only on the dimensionless coordinates 2/d, y/d, z/d. If, however, the two systems are geometrically, but not dynamically, similar, i. 0. if their Reynolds nombers are different, then the dimensionless quantities under consideration must also dopend on the characteristic quantitics V,d, @, 2 of the two systems. Applying the principle that. physical laws must he independent of the system of units, it follows that the dimensionless quantities u/V,..., pio V?,t/@V? can only depend on a dimensionless combination of V,d, 9, and y which is unique, being the Reynolds number R = V d g/. Thus we are led to the conclusion that for the two geometrically similar systems which have different Reynolds nunihers and which arc being conipared, the dimensionless quantities of the field of flow can only be functions of the three dimensionless space coordinates x/d, y/d, 2/d and of the Reynolds number R. ‘The preceding dimensional analysis can be atilized to make an important assertion about the total fore exerted by a flnid stream on an immersed body. The force acting on the body is the surface integral of all normal aud. shearing stresses acting on it. If P denotes the component of the resultant foree in any given direction, ib is possible to write a dimensionless force coefficient of the form P{d? 9 V2, but in- stead of the area d? it is customary to choose a differcut characteristic arca, A, of the immersed body, e.g. the frontal aren exposed hy the body to the flow direction which is, in the case of a sphere, equal to x d?/4. Hence the dimensionless force coef- ficient becomes P/A 9 V2, Dimensional analysis leads to the conchision that for geometrically similar systems this coefficient. can depend only on the dimensionless gronp formed with V,d,@, and p, i.e. on the Reynolds number. ‘The component 16 J. Outline of fluid motion with friction of the resultant force parallel to the undisturbed initial velocity is referred to as the drag D, and the component perpendicular to that direction is called lift, Z. Hence the dimensionless coefficients for lift and drag become D and Cy = 55 if the dynamic head } 9 V? is 5 selected for reference instead of the quantity @ V?. ‘Thus the argument leads to the conclusion that the dimensionless lift and drag coofficients for geometrically similar systems, i. o. for geomnetrically similar bodies which have the same orientation with respect to the frec-stream direction, are functions of one variable only, namely the Reynolds number C, =f(R); Cy = fa(R)- (1.15) Iis necessary to stress once more that this important conclusion from Reynolds's principle of similarity is valid only if the assumptions underlying it are satisfied, ie. if the forces acting in the flow are due to friction and incrtia only. In the case of compressible fluids, when elastic forees are important, and for motions with free surfaces, when gravitational forces must be taken into consideration, eqns. (1.15) do not apply. In such cases it is necessiry Lo deduce different, similarity principles in which the dimensionless Froude number F = V//gd (for gravity and incrtia) and the dimensionless Mach number M = V/e (for compressible flows) are included. ; (1.14) ‘The importance of the similarity principle given in eqns. (1.14) and (1.15) is very great as far as the sciences of theoretical and experimental fluid mechanics are concerned. First, the dimensionless cocfficients, C,, Cp, and R are independent of tho system of units. Secondly, their usc leads to a considerable simplification in the extent of experimental work. In most cases it is impossible to determine the functions {,(R) and /,(R) theoretically, and experimental methods must be uscd. d Supposing that it is desired to determine the drag cocfficiont, Cp for a sp shape of holy, e. g.asphere, then withont the application of the principle of simila it would be necessary to carry out drag measurements for four independent variables, V,d,e, and g, and this would constitute a tremendous programme of work. It follows, however, that the drag coefficient for spheres of different. diameters with different stream velocities and different fluids depends solely on one variable, the Reynolds number. Fig. 1.4 represents the drag coofficient of circular cylinders as a function of the Reynolds number and shows the execllent. agreement. between experiment. and Reynolds's principle of similarity. The experimental points for the drag coefficient. of circular eylindors of widely differitig diameters fall on a single curve. The same applies to points obtained for the drag coefficiont of spheres plotted against the Reynolds number in Fig. 1.5. The sudden decrease in the value of the drag coefficient which occurs near R = 5 x 10% in the case of cireular cylinders and near R = 3 x 105 in the case of spheres will be discussed, in more detail, later, Fig. 1.6 reproduces photographs of the streamlines about circular cylinders in oil taken by I’. Homann [7]. They give a good idea of the changes in the field of flow associated with various Reynolds numbers. For small Reynolds numbers the wake is laminar, but at increasing Reynolds numbers at first very regular vortex patterns, known as Kfrmfn’s vortex streets, are formed. At still higher Reynolds numbers, not shown here, tho vortex patterns become irregular and turbulent in character. c, Principle of similarity; Lhe Reynolds aud Mach numbers 17 TH-HHE = fn 08) or a3 10 | tasaee in| 225 [oesettorgr 2 azo 3 see 3 000 Then te tomb oeeeo Lt. 1? a 52 > a 2 68 4 o 6 pit PONGO TE OMG TF Ogg? FONG PH EAL Te OR GE FE ARs z% Fig. 1.4. Drag coefficient for circular cylinders a8 a function of the Reynolds number Measured by: © Schiller = Schmecet © Liebster © Allen ' © 1924 2 1976 }mesetsberger | T|EAye tr pe? PBps 2 + 68 Re gt? * 882? 4 6A s? 4 68, Fig. 1.5. Drag coefficient for spheres asa function of the Reynolds number Curve (1): Stokes’s theory, eqn. (6.10); outve (2): Oseen’s theory, eqn. (6.13) 13 T. Onttine of Haid motion with friction R= 102 R ~ 161 R — 295 R-: 281 Fig. 1.6. Field of flow of off about a cirentar cylinder at varying Reynolds numbers after Homann 17]: transition from laminar How to a vortex strect in laminar flow. The frequency range for R= 65 to R = 281 can be taken from Mig, 2.9 In Inter times such pictures for low Reynolds numbers up to R= 3 were produced by 8. 'Tanoda [14]. It is seen in Wig. 1.4 that the drag coefficient of a circular cylinder reaches @ minimum of Cp © 0.3 at- a Reynolds number between R = 5 x 105 and 108, A regular vortex strect «loes not exist in this range of Reynolds numbers. At very high Reynolds numbers exceeding R © 10%, the drag coefficient. inereases at a considerable rate, as scen from Fig. 1.7 which is based on the measnrements performed by A. Roshko {13} and GW. Jones, J.J. Cinotta and R, W. Walker [8]. At R= 10? the drag cocfticient, reaches a value of Ep ~ 055. According to the preceding anthors, a regular vortex strcet establishes itself again at R > 35 x 108, The drag of spheres has recently also heen investigated at. very high Reynolds numbers [1]. Here loo, as was the case with the cylinder, the drag coclficient inereases appreciably beyond its minimum at Cp © O-L at about R = 5 x 108 attaining Cp © 0-2 at. Reynolds mimbers close to R = 107. Critical reviews of drag measurements on spheres as a function of the Reynolds number and the Mach number were prepared by A. B. Bailey and J. Tiatt [la] as well ag by A.B, Bailey and RB. F, Starr [1b]. e. Principle of similarity; the Reynolds and Mach numbers: 19 M12 20 30 ts 10 a 08 a7: 06 0 Fiacheebart (1982) 908 (2 Roshke (1961) eo HF Et © NACA (1969) Rey Fig. 1.7. Drag cocfficiont of x circular cy. der at very large Reynolds numbers and for Mach nombers M 0-2 after the mensnrements of A. Roshko [13] and G. W. Jones, J. J. Cinotta and 1, W, Wat- ker (8) Fig. 1.8. Drag coefficient of sphe of the Reynolds and Mach number red by A. Nauman [, 10] es in terms. Influence of compressibility: The preceding argument was conducted under the assumption that the fluid was incompressible, and if was found that the dimensionless depeucent quantities were finctions of one dimensionless arguincnt, the Reynolds uuinber, only. When the fluid is compressible they depend ou an additional dimeusion- less munber, the Mach number M = V/e which can he regarded, a8 shown in See. Te, as a measure of the compressibility of the flowing medium. In the case of suctt flows, i.e, when conipressibility plays an essential part, Ue dimensionless coefficients dopend on both parameter Rand M. Equation (1.15) is then ropliced hy Cy -= (RM); Cy = fy (R,M). (1.16) An example of sich a relationship is given in Fig, 1.8 which shows a plot of the drag cocflicient Cy of spheres in ternis of the Reynolds mumber Ri VF Dm aud the Mach number M = F/e. The curve for M = 0-3 is practically coincident with that in Fig. 1.5 for incompressible low which proves that upto M == 0-3 the influence of the Mach number is negligible. On the other hand at higher Mach numbers the influence is large. In this connexion it is noteworthy thal, in the range of Reynolds numbers covered by the diagram, its mflncnce reecdea more and more as the Mach number is increased, 20 T. Outtine of fluid motion with friction f. Comparison between the theory of perfect fl Js and experiment In the cases of the motion of water aut air, which are the most important ones in engineering applications, the Reynolds numbers are very large because of the very low viscosities of these fluids. It would, therefore, appear reasonable to expeck very good agreement between experiment. and a theory in which the influence of viscosity is neglected altogether, i. c. with the theory of perfect fluids. In any ease it seems uscful to begin the comparison with experiment by reference to the theory of perfect fluids, if only on account of the large number of existing explicit. mathe- matical solutions. In fact, for cortain classes of problems, sel as wave formation and tidal motion, excellent. results were obtained with the aid of this theoryf. Most problems to be discussed in Uris book consist in te study of the motion of solid bodies through fluids at rest, or of fluids flowing through pipes and channels. In such eases the use of the theory of perfect fluids is limited because its solutions do, not satisfy the con- dition of no slip at the solid surface which is always the case/with real fluids even at very sinall viscosities. In a perfeet fluid there is slip at a qwall, and this cireum- stance introduces, even for smull viscosities, such fanddamoyial differences that it is rather surprising to find in some cases (c. g. in the case of very slender, stream-line hodies) that the two solutions display a good measure of agreement. ‘The greatest diserepaney hetween the theory of a perfect fluid and experiment exists in the consideration of drag. The perfeet-fluid theory leads to the conclusion that when an arbitrary solid body moves through an infinitely extended fhrid at rest it ex- periences no foree acting in the direction of motion, i. ¢. that its drag is zero (d’ Alem- bert’s paradox). This result is in glaring contradiction to observed fact, as drag is measured on all bodies, even if it can become very small in the case of a stream- line body in steady flow parallel to its axis. By way of illustration we now propose to make some remarks concerning the flow about a circular cylinder. The arrangement of streamlines for a perfect. fluid is given in Fig. 1.9. It follows at once from considerations of symmetry that the resultant. force in the direction of motion (drag) is equal to zero. The pressure distribution according to the theory of frictionless motion is given in Fig. 1.10, together with the remilts of measurements at Uhree values of the Reynolds number. At the leading edgo, all measured pressure distributions agree, to a certain extent, with that for a perfect Muid. At the trailing end, the discrepancy between theory aud measurement becomes large because of the large drag of a circular cylinder. ‘The pressure diatri- bution at the lowest, subcritical Reynolds number R = 1-9 x 10° differs most. from that given by potential theory, The measurements corresponding to the two largest Reynolds numbers, R = 67 x 105 and R = 8-4 x 105, are closer to the potential curve than those performed at the lowest Reynolds number. The large variation of pressure distribution with Reynolds number will be discussed in detail in the next chapter. A corresponding pressure-distribution curve around a meridian section of a sphere is reproduced in Fig. 1,11, Hore, too, measurements show large differences for the two Reynolds numbers, and, again, the smaller Reynolds number lies in the range fC eg BLU. Tanth: Mydradynamies, 6th od., Dover, New Vork, 1945, Ff. Comparison between the theory of perfect fhtids and experiment 21 Fig. 1.9. Frictiouless flow about a circular cylinder Fig. 1.10, Pressure distribution on a circular cylinder in the subcritical and supercritical range of Reynolds numbers after the measurements of 0. Flachsbart [4] and A. Roshko [13}-9o = 9 V2 is the stagnation pressure of the oncoming flows frlctioniess Low plachabart Roa 4932) Rorbko (1061) Te Fig. 1.11. Pressure distribution Suberiicat sround a sphere in the suberi- ‘Supercritical tical and supercritical range of 9} |__|}. Reynolds numbers, as_mea- sured by O. Flachsbart (3] of large drag coefficients, whereas the larger value lids in the range of small ding coefficients, Fig. 1.5. In this ease the mcasnred pressure-distribution curve for the large Reynolds number approximates the theoretical durve of frictionless flow very well over the greatest, part of the circumference. Considerably better agreement between the theoretical and measured pressure distribution is obtained for a streamline body in a flow parallel to its axis [5], Fig. 1.12. Good agreement exists here over almost the whole length of the body, with the exception of a small region near its trailing end, As will be shown later this circumstance is a consequence of the gradual pressure inereasé in the down- stream direction. Although, generally speaking, the theory of perfect fluids docs not lead to useful results as far as drag calculations are concerned, tlie lift can be calculated from it very successfully. Fig. 1.13 represents the relation between the lift coefficient aud augle of incidence, as ieasured by A. Betz [2] in the case of a Zhukovskii acrofoil 22 1. Outline of fluid motion with friction of infinite span and provides a comparison with theory. ln the range of incitlence angles a =: — 10° to 10° the agreement is seen to be goo and the small differences can be explaineil by the influence of friction. ‘The measured and calculated pressure distributions agree very well tov, as shown in Wig, 1.14. The discrepancy between theory and measurement displayed in Figs. 1.13 and 1.14 is a consequence of the displacement. action of the boundary layer and conatitutes a boundary-layer effect of higher order, as will be shown again in See. LX}. 10 Fig. 1.12, Pressure distribution about a streanvtine body of revolution; comparison —bet- ween theory and measurement, after Fntrmann (5] Fig. 1.13. Lift and drag coeffi- ciont of a Zhukovskil profik plane flow, as measured by Betz {2} References Fig. 1.14. Comparison between the theoreticnt and measnred prensure distribul for a ‘Alnkovstii profile at equal Bifts, after A. Bots {2} References Y Achentart, APM 5d, 575 (1972) [a] Baitey, A. "Bosna dat. nmmbers. ATAA J, 10, 1430 ~ 1440 (1972), [1b] Bailey, A.B, and Starr, RUF, : Sphere drag at trangonic speeds and high Reynolds nnmbers. AIAA'I. 14, 1631 (1976) {2] Betz, A.: Untersuchun (3) Flachsbart, 0.: Neuere 461 469 (1927). 14] Machsbart, 0.: Winddr 134—138 (1932), (5) Fobrmaun fi. Wine Coitti Experiments on the flow past spheres at very high Roynolds nnnsbers i Sphere drag coefficients fora broad range of Much and Reynolds nor Foukowskischen ‘Tragtliiche, ZM 6,173 179 (1915). ersnclungen fiber den Luflviderstand von Kugein. Phys. Z. anf Gosbehalter. Reporte of the AVA in Géttingen, IVth Series, ‘aeoretische und experimentelle Unterachangen an Ballonmadelten. Ab. Motortaftschill-Stndienges. V 63-- 123 (1911/12). vegimg cles Wassera in engen zylindrisehen Réhren. Pogg. Aun. 16, 423—442 (1 17] Homann, P. 1-- 10 (1936), [8] Jones. G.W., Cinotta crodynainic forces on a stationary and mbers. NACA TR R-300 (1969) [9] Naumann, A.: Luftwiderstand von Kogetn bei hohen Unterschaligerchwindigkciten. All. gem. Walrinetochnite 21 (19533 [10) Naumann, A. Sher tie italrdimang sin Zylinder bei loher eschwindigkeiter, Advances in Acronautical Sciences (Tt. von Kéranin, ed.) Vol, 3, 185-206, London, 1952. [11] Poiscnitie, J.: Récherches expérimentettos sur te mou yentent. des tiquides dans tos tubes de trés pr tron, Comptes Rendus 11, 961907 and 1041-1048 (1840); 12, 12-115, (1841); in more detail: Mémoires des Savanta Etrangors 9 (1840). [12] Reynolds, O.: An experimental investigation of the cireumstances which deter the motion of water shall be direct or simous, and of the taw of resistance in paralte Phil. ‘Trans, Roy, Soc. 171, 915—982 (1883) or Scientific Papers UL, BI. 113] Roshko, A. Experiments on the flow past n cireular cylinder at very high Reynolds numbers, JEM 70, 345-356 (1961); sec also: On the acrodynamic drag of eytinters at high Reynolds nunbers. Paper presented at the US Japan Research Sominar ot Wind Loads on Strachures, Uniy. of Hawaii, Oct. 1970, [14] Taneda, 8.: Experimental investigation of the wakes heltind cylinders and plates at low Reynolds mumbers. J. Phys. Soc. Japan 17, 302-307 (1956). ne whether cliammets. CHAPTER It Outline of boundary-layer theory a. The boundary-laycr concept In the case of fluid motions for which the measured pressure distribution nearly agrees with the perfect-fluid theory, such as the flow past the streamline body in Fig. 1.12, or the acrofoil in Fig. 1.14, the influence of viscosity at high Reynolds numbers is confined to a very thin layer in the immediate neighbourhood of the solid wall. If the condition of no slip were not to be satisfied in the case of a real fluid there would be no appreciable difference between the field of flow of the real fluid as compared with that of a perfect fluid. The fact that at the wall the fluid adheres to it means, however, that frictional forces retard the motion of the fluid in a thin layer near the wall. In that thin layer the velocity of the fluid increases from zero at the wall (no slip) to its full valuc which corresponds to external frictionless flow. The layer under consideration is called the boundary layer, and the concept is due to L, Prandtl (25). Fig. 2.1, Motion along thin flat plate, from Prandtt-Tietjene J = length of plates Reynolds number R= Hiv = 3 Fignre 2.1 reproduces a picture of the motion of water along a thin flat plate in which the streamlines were made visible by the sprinkling of particles on the surface of the water. The traces left by the particles arc proportional to the velocity of flow. It is scen that there is a very thin layer near the wall in which the velocity is’ considerably smaller than at a larger distance from it. The thickness of this bonndary layer inercases along the plate in a downstream direction. Fig. 2.2 repre- sents diagrammatically the velocity distribution in such a boundary layer at the @, The boundary-layer concept. 25 plate, with the dimensions across it considerably exaggerated. In front of the loading edge of the plate the velocity distribution is uniform. With increasing distance from the leading edge in the downstream dirootion the thickness, d, of the retarded layor increases continuously, as inercasing quantitics of fluid become affertod, Evidently the thickness of the boundary layer decreases with ilecreasing viscosity. Fig. 2.2. Sketch of boundary layer on a flat plate in par- allel flow at zero incidence On the other hand, even with very small viscosities (large Reynolds numbers) the frictional shearing stresses r = x Ou/dy in the boundary layer are consilerable because of the large velocity gradient across the flow, wherens outsile tho boundary layer they are very small, This physical picture suggests that the field of flow in tho case of fluids of small viscosity can be divided, for the purpose of mathematical analysis, into two regions: the thin boundary layer near the wall, in which friction must be taken into account, and the region outside the boundary layer, where the forces due to friction are small and may be neglected, and where, therefore, the perfect-fluid theory offers a very good approximation. Such a division of the field of flow, as we shall see in more detail later, brings about a considerable simplification of the mathematical theory of the inotion of fluids of low viscosity. In fact, the theoretical study of such motions was only made possible by Prandtl when he introduced this concept. We now propose to explain the basic concepts of boundary-layer theory with the aid of purely physical ideas and without the use of mathematies. The mathemati- cal boundary-layer theory whieh forms the main topic of this book will be discussed in the following chapters. ‘The decelerated fluid particles in the boundary layer do not, in all cases, remain in the thin layer which adheres to the body along the whole wetted length of the wall. In some cases the boundary layer increases its thickness considerably in the downstreain direction and the flow in the boundary layer becomes reversed. ‘This causes the decelerated fluid particles to be forced outwards, which means that the boundary layer is scparated from the wall. We then speak of boundary-layer separation. This phenomenon is always associated with the formation of vortices and with large energy losses in the wake of the body. It occurs primarily near blunt bodies, such as circular cylinders and spheres, Behind such a body there exists a region of strongly decelerated flow (so-calied wake), in which the pressure distribution deviates considerably from that in a frictionless fluid, as seen from Figs.1.10 and 1.11 in the respective eases of a cylinder and a sphere. The large drag of such bodies can be explained by the existence of this large deviation in pressure distribution, which is, in turn, a consequence of boundary-layer separation. 26 II. Outtine of boundary-layer theory Estimation of boundary-layer thickness: The Usickness ofa boundary layer which has not separatetl can be casily celimated in the following way. Whereas friction forees ean be neglected with respect to inertia forees outside the boundary layer, owing to low viscosity, they are of a comparable order of magnitude inside it. ‘The inertia force per unit. volume is, as explained in Section Le, equal to 9 udu/dz. For a plate of length U the gradient aujéx is proportional to U/l, where U denotes the velocity outside the bountary layer. Uenee the inertia force is of the order g U/l. On the other hand the friction foree per unit volume is equal to dr/ay, which, on the assumption of laminar flow, is equal to 4 2%u/dy2. The velocity gradient aujay in a direction perpentienlar to the wall is of the order U/5 so that the friction foroe per unit volume is dr/oy ~ ye 1/6. From the condition of equality of the friction and inertia forces the following relation is obtained: U M §8 7 or, solving for the boundary-layer thickness 5+: i vt anna a . (2) The numerical factor which is, so far, still undetermined will be deduced later (Chap. VIX) from the exact solution given by U. Blasins [4], and it will turn ont that it is eypial Lo 5, approximately. Hence for laminar flow in the boundary layer we have b= 5 Vi : (2.18) ‘The dimensionless boundary-layer thickness, referred to the length of the plate, L. becomes: = vy =5V i= whore R, denotes the Reynolds number related to the length of the plate, l. Tt is acon from eqn. (2.1) that the boundary-layer thickness is proportional to ¥ » and to VL. If Lis replacetl by the variable distance x from tho leading edge of the plate, it is seen that J increases proportionately to Vz. On the other hand the relative boundary-layer thickness 5/l decreases with increasing Reynolds mmber as 1///R stionless flow, with R~> oo, the boundary-layer (2.2) so that in the limiting case of fr thickness vanishes. We are now in a position to estimate the shearing stress tr, on the wall, and consequently, the total drag. According to Newton’s law of friction (1.2) we have a a=n(%),, A more rigorous definition of bountlary-layer thickness is given at the end of this section. a. The boundary-layer concept 27 where subscript 0 denotes the value at the wall, i.e. for y = 0. With the estimate (du/Ay)) ~ U/6 we obtain ty ~ # U/S and, inserting the value of 5 from eqn. (2.1), we have (2.3) ‘Thus the frictional stress near the wall is proportional io U3, We can now form a dimensionless stress with referenee to @ U2, ax expluinedl in Chap. I, and obtain Te ” ae~ (2.32) This result agrecs with the dimensional analysis in Chap. 1, which predicted that the dimensionless shearing stress coukl depend on the Reynokls number only. The tatal drag 2) on the plate is equal to bly where b denotes the width of the plate. Hence, with the aid of eqn. (2.3) we obtain DrbYonWl. (2.4) The laminar frictional drag is Unus seen to be proportional to ("2 and 1, Pro. portionality to U2 means that doubling the plate length docs not denble the dag, and this result. can be understood by considering that the downstream part. of the plate experiences a smaller drag than the leading portion because the boundary layer is thicker towards the trailing edge. Finally, we can write down an expression for the dimensionless drag coefficient in accordance with eqn. (1.14) in whieh the reference area A will be replaced by the wetted area bl. Hence eqn, (2.4) gives that, The numerical factor follows from IT. Blasius's exact solution, and is 1-328, so that the drag of a plate in parallel laminar flow becomes The following numerical example will serve to illust coctling estinia tion : Laninar flow, stipulated here, is oblained, as is known from experiment, for Rey- nohls numbers Ulfy not exceeding ahout 5 x 105 to 10% For larger Reynolels numbers tire boundary layer becomes turbrilent. We shall now calenlate the houndary- layer thickness for the flow of air (v = 0-144 x 10-* ft/sec) at the end of a plate of fength [= 3 ft at a velocity U = 48 ft/sec. This gives Ry = Ul/y ~~ 10" and from eqn. (2.2) = 0005; b= O1Rin, The drag coefficient from eqn. (2.5) is Cy = 0-013 i. ¢. exceedingly small when compared with that for a circular cylinder, Fig. 1.4, because the drag coefficient. far a. eylinder also inelutles pressnre forces. 28 11, Outline of boundary-layer thoory Definition of boundary-layer thickness: The definition of the boundary-layer thickness is to a certain extont arbitrary because transition from the velocity in the boundary to that. outside it takes place asymptotically. ‘This is, however, of no practical importance, because the velocity in the boundary layer attains 2 value which is very close to Ute external velocity already at a small distance from the wall. Ft is possible to define the boundary-layer thicknoss as that distance from the wall where the velocity ditfers by | per cent from the external velocity. With this definition the numerical factor in eqn, (2.2) las the value 5. Instead of the boundary- layer thickness, another quantity, the displacement thickness 54, is sometimes uscd, Mig. 2.3. 1b is defined by the equation vo, f(U~way. (2.6) ‘ Fig. 2.3. Displnecment thickness 6, in a boundary layer The displaccmnent thickness indicates Ue distance by which the external stream- lines are shifted owing to the formation of the houndary layer. In the case of a plate in parallel flow and al zero incidence the displacement. thickness is about } of the boundary-layer thickness 5 given in eqn. (2.19). b, Separation and vortex formation The boundary layer near a flat plate in parallel flow and at zero incidence is particularly simple, because the static pressure remains constant in the whole field of flow. Since outsile the boundary layer the velocity reniains constant the sane applies ty the pressure because in the frictionless flow Bernoulli's equation remains valid. Furthermore, the pressure remains sensibly constant over the width of the hountlary layer at a given distance x, Hence the pressure over the width of the boundary layer has the same magnitude as outside the boundary layer at the same distance, and the sare applics to eases of arbitrary body shapes when the pressure outside the boundary layer varies along the wall with the length of are. ‘This fact is expressed by saying that the external pressure is “impressed” on the boundary layer, Hence in the case of the motion past a plate the pressure remains constant Uronghout the bountary layer. The phenomenon of boundary-layer sepa ranynenine previouslyisintimately connected with the pressure distribution in U6 baiindary layer. In the boundary layer on a plate no separation takes place as no back-f osétirs. + In order to explain the very important phenomenon of boundary-layer separation lot. us consider the flow about a blunt. body, ¢. g. abouta circular cylinder, as shown in Fig, 2.4, In frictionless flow, the fluid particles are accelerated on the upstream b. Separation and vortex formation 29 half from D to £, and decelerated on the downstream half from # to F. Hence the pressure decreases from D to # and increases from E to F. When the flow is started up the motion in the first instant is very nearly frictionless, and remains so as long as the boundary layer remains thin, Outside the boundary layer there is a tranafornmtion of pressure into kincti gy long DB, the reverso taking placn along Hi (hat a particle arrives at, # with the samo velocity ns it lind at D. A fluid parti which moves in the immediate vicinity of the wall in the boundary layer remains under the influence of the same pressure field as that oxisting outside, because the external pressure is impressed on the boundary layer, Owing to the large friction forces in the thin boundary layer such a particle consumes so much of its kinktic Fig. 2.4. Boundary-layer separa- tion and vortex formation on a circular cylinier (diagrammatic) 8 = point of separation energy on its path from D to # that the remainder is too sinall to surmount the “pressure hill” from # to F, Such a particle cannot move far into the region of increasing pressure between # and F and its motion is, eventually, arrested. ‘The external pressure causes it then to move in the opposite direction. The photographs reproduced in Fig. 2.5 illustrate the sequence of events near the downstreain side of a round body when a fluid flow is started. The pressure increases along the body contour front left to right, the flow having been made visible by sprinkling aluminium. dust on the surface of the water. The boundary layer ean be easily recognized by reference to the short traces. In Fig. 2.5a, taken shortly after the start. of the motion, the reverse motion has just begun. In Fig. 2.5b the reverse motion has ponztrated a considerable distance forward and the boundary layer has thickened appreciably. Fig. 2.5 ¢ shows how this reverse motion gives rise to a vorlex, whose size is increasod still further in Fig. 2.5d. ‘The vortex becomes separated shortly afterwards and moves downstream in tho fluid, This cirenmstance changes completely the field of flow in the wake, and the pressure distribution suffers a radical change, as compared with frictionless flow. The final state of motion can be inferred from Vig. 2.6. In the eddying region behind the cylinder there is considerable suction, as seen fron: the pressure distribution curve in Fig. 1.10. This suction causes a large pressure drag on the body. : At a larger distance from the body it is possible to discern a regular patte of vortiecs which move alternately clockwise and counterclockwise, aud which known ag a Karman vortex strect [20], Fig. 2.7 (sce also Fig. 1.6). In Fig. 2.6.a vortex moving in a clockwise direction can be scen to be about to detach itself from the body before joining the pattern. In a further paper, von Kérmén [21] proved that such vortices are gencrally unstable with respect to small disturbances parallel 30 I. Outline of boundary-layer theory ig. 2. Fi oF separation with time, after Prandtl-Tietjons 127). b,c. Development, of boundary ye also 15.5 to themselves. The only arrangement which shows nentral equilibrium is that with Aft -= 0-281 (Vig. 2.8). The vortex street moves with a velocity %, which is smaller than the flow velocity U in front of the body. It can be regarded as a highly idealized picture of the motion in the wake of the body. The kinetic energy contained in the velocity ficld of the vortex street must be continually created, as the body moves throngh the fluid. On the basis of this representation it is possible to deduce an expression for the drag from the perfect-fluid theory. Its magnitude per unit length of the cylindrical body is given by 2 D= oUt 283 + —1ie(Hy]. A "The width h, aul the velocity ratio u/{7 must be known from experiment. More ree that in an av vortices changes cow a turbulent wake. perimental investigations due to W. W. Durgin and others [13] established clerating vortex sirvet the ratio of the fon, al to the transverse spacing of the erably. Asa result, the regitar arrangemeut. of vortives is transformed into, b. Separation and vortex formation 31 Fig. 2.6 Fig. 2.6. Jnstantancons photograph of flow with complete honndary-layer separation in the wake of a circular cylinder, after Prandul-Tietjous {27} Fig. 2.7. Kérmén vortex street, from A. Timme [38] Fig. 2.8. Streamlinesin a vortex street. {h/l = 0-28), The fluid is at reat at infinity, and the vortex street moves to the left, Fig. 2.8 Cireular cylinder. The frequeney with which vortices arc shed ina Karmén vortex street behind a cireular cylinder was first extensively measured by I. Blenk, D. Firchs and L. Licbors [5]. A regular Kérmén street is observed only im the range of Reynolds numbers V D/v from about 60 to 5000. At lower Reynolds numbers the wako is laminar and has the form visible in the first two photographs of Vig. 1.6; at higher Reynolds nuntbers there is complete turbulent mixing, Measurements show that in the regular range given above, the dimensionless frequcney, "28, — (Strouhal number) also known as the Strouhal number [37], depends only on the Reynolds wun ter. ‘This relationship is shown plotted in Fig. 2.9 which is based on measurements per- formed hy A. Rashko [32|; see also [15]. The experimental points which were ob- tained with eylindors of different diameters D and at. diffvrent velovitivs V arrauge themselves well on a single curve. At the higher Reynolds numbers the Stvonlial number remains approximately constant at $ = 0-21. This value of S, as seen front Fig. 2.9, prevails up to a Reynolds number R = 2 x 105, that. is in the snberitical range (see also Fig. 1.4). At higher Reynolds numbers, say around R= 108, a regular vortex strect: does not exist. Accordiug to A. Roshko |31}, such a regular strect re-appears at. extremely large Reynolds mmmbers (R > 3 x 105) when the Stroulal 32 IL. Outline of boundary-layer theory Vv D= 3+ 100m Frinberger () Dw 24 1B em Hig. 2.9, The Strouhal number, S, for the Karman vortex strect in the flow past a cirentar cylinder in terms of the Reynolds number, R. Measurements performed by A. Roshko (31. 32), IL. 8. Ribner, 1B, Btkins and K. K. Nelly (30), B. ¥. Relfand L, F. G. Simmons [28] as well as G. W. Jones ot al. ((8] of Chap. 1). In the range R = 3 x LO to 3 x 108 (supercritical regime with very low dmg. Fig 1.4) the Kérm&n vortex strect is no loger regular. It is only at R > 4 x 10% that a regular pattern forme again; its Strouhal nuinber is now higher at S = 0-26 to 0-30 compared with S 0-2 at R= 103 (03 x 105 mumber assumes values around $ = 0-27. In this connexion the paper by P.W. Bear- maui [3a] may also be consulted, When the diameters of the cylinders are small and the velocities arc moderate, the resulting frequencies lie in the acoustic range. For example, the fainiliar “aeolian tones” emited by telegraph wires are the result of these plicnomena. At a velocity of ¥ = 10 m/sec (30-48 ft/sec) and a wire of 2mm (0-079 in) in diaineter, the frequency becomes n = 0-21 (10/0-002) = 1050 sec-!, and the corresponding Reynolds number R ~ 1200. Flat plate at zero incidence. The fact that a regular vortex street establishes it- self, among others, beliind slender bodies as well as in compressible streams has only been established recently by H. J. Heinemann ct al. [18]. The photograph of Fig. 2.10 shows such a regular vortex street behind a flat plate at zero incidence for a Machi number Mo = 0-61. The diagram in Fig. 2.11 contains a plot of the Strouhal number, S = nd/V, formed with the plate thickness, d, in terms of the Mach number, but only for Ute subsonic range M = 0-2 to 0-85. The diagram proves that liere too § ~ 0-2, as was the case for the circular cylinder in Fig. 2.9. The corresponding Reytiotds numbers, referred to the length of the plate, are in the range R = Vly = 3 x 105 to 8 x 105 in which the flow is laminar. ' Two papers by C. C. Lin. [22] and U. Domm [II] concern themselves with the theory of the Karman vortex street, The formation of a vortex pair behind a flat plate in cross-flow at right angles to it has been investigated theoretically by E. Wede- uicyer [38a], whereas ‘I, Sarpkaya [33h] conducted theoretical and experimental studies fora plate arranged ata large angle of attack (sce Fig. 4.2); in this connexion b. Separation and vortex formation 33 Fig. 2.10, Von Kériniin vortox atreot behind Fig. 2.1L Strouhal number Sd] in a flat plate at zero incidence at a Mach num- terms of the Mach nuinber for the vortex ber M = 0:61 and a Reynolds number R= street behind a flat plate at zero incidence, Viv = 6-6 x 105 after H. J. Heinemam ct after H. J. Heinemann et al. (18) al, [18]. Length of ate L = 60 mm, thickness ratio d/l = 0-05, Exposure time approx. 20 rianosee (20 x 10° sec) an earlier paper by L. Rosenhead [32a] may also be consulted. The reader may also be interested to look up the text of a remark made by LL. Prandtl on the occasion of a lecture by K. Friedrichs (,,Bemerkung tiber die ideale Strémung uin einen K6rper bei verschwindender Zahigkeit‘' Lectures on aerodynamics and allied subjects, Aachen 1929, Springer, Berlin 1930, pp. 51, 52). Scparation. Tie boundary-layer theory succecds in this manner, i.e. with the aid of the explanation of the phenomenon of separation; in throwing light on the occur- renee of pressure or form drag in addition to viscous drag. ‘The datigor of boundary- layer separation exists always in regions with an adverse pressure gradient and tite likelihood of its occurrence increases in the ease of steep pressure curves, i.e, behind bodies with blunt ends. The preceding argument explains also why the experimental pressure distribution shown in Fig. 1.11 for the case of a slender streamline body differs so little from that predicted for frictionless flow. The pressure increase in the downstream direction is here so gradual, that there is rio separation, Consequently, there is no appreciable pressure drag and the total drag consists mainly of viscous drag and is, therefore, small. ‘The streamlines in the boundary layer near separation are shown diagrammatic- ally in Fig. 2.12, Owing to the reversal of the flow there is a considerable thickening of the boundary layer, and associated with it, there is a flow of boundary-layer material into the outside region, At the point of separation onc streamline inter- 11. Quttine of bonndary-layer theory 5 = point of separation Fig. 2.12, Diagrammatic reprosen- tation of flow in the boundary layer near a point of separation Fig. 2.13. Flow with separation in a highly divergent channel, from Prandtl-Tietjens [27] Ps Fig. 2.14. Flow with boundary- layor suction on upper wall of highly divergent: channel Fig. 2.15. Flow with boundary- layer suction on both walls of highly divergent channel b. Soparation and vortex formation 30 scols the wall at a definite angle, and the point of separation itself is determined by the condition that the velocity gradient normal to the wall vanishes ther ‘The precise location of te point of sepa of an exact calomlation, i.e. by the ir tion can be determined only with the aid ution of the boundary-layer equations. Separation, as deseribed for the case of a circular cylinder, ean also occur in a highly divergeut channel, Fig. 2.13. In front of the throat the pressure decreases in the direction of flow, aud tte flow ailheres conipletely to the walls, as itt a frictionless fluid. However, behind tho tiroat the divergetice of the channel is so large that the boundary layer becomes separated from bolli walls, and vortices are formed. The stream fills now only a sinall portion of the cross-sectioual arca of the chanucl. How- ever, separation is prevented if boundary-layer suction is applied at. the wall (Figs. 2.14 and 2.15). The photographs in Figs. 2.16 and 217+ prove that the adverse pressure gradiont together with friction near the wall determine the pracess of separation which is independent of such otfier circumstance as c. g. the curvature of the wall. ‘The first picture shows the motion of a fluid against. a wall at right angles to it (plane stagnation flow), Along tlie streamline int the plane of symmetry which leads to the stagnation point there is a cotisiderable pressure increase in te direction of flow. No separation, however, occurs, becanse no wall friction is present. ‘There is no separation near Ute wall, cither, because here the flow in the boundary layer takes place in the direction of decreasing pressure on both sides of the plane of symmetry. Tf now a thin wall is placed aloug the plane of symmetry at right angles to the first. wall, Fig. 2.17, the new boundary layer will show a pressure increase i the dircetion of flow. Consequently, separation now oceurs near the plane wall, The incidence of se is often rather sensitive to small changes in the shape of the solid body, par when the pressure distribution is strongly affected by this change in shape. A very instructive example is given iu the pictures of Fig. 2.18 which show photograplis of the flow field about a model of a motor velticle (the Volkswagen delivery van), [23, 35]. When the nose was flat. giving it an angular shape (a), the flow past the fairly sharp corucra in frout caused large suction followed by a large pressure increase along the side walls, This led to coniplete separation aud to the formation of a wide wake behind the body. The drag coefficient of Ute velticle with this angular shape had a value of Cp = 0-76. The large suction near the front end and the separation along the side walls were eliminated when the shape was changed by adding the round nose shown at (b). Simultancously, the drag eoefitciet beeame imarkeclly smaller and had a value of Cp = 0-42. Further research on such vehicles have beet performed by W. H. Helo [19] for the case of a non-syinmetric stream. ig. 2.16. and 2.17. have been taken from the paper “Strému by H. Foottinger, Mittcitungen der Vereinignnd A& AHS: Kesselbesitzer, No. eu in Dampfikcssclantagen” 3, pp. GL (1939). 36 Th. Fig. 2.16. Free stagnation flow withont sepa- ration, as photographed by Focttinger (a) Angular nose | | | a9 2.18. Flow about a model of « motor vehicle (Volkawagen delivery van), after B. Moeller a) Angular tose with separated flow along the whole of the side wall and large drag eoef- ficfent (Cp = 0-76); b) Round nose with no separation and small drag coefficient (Cp = 0-42) 'b. Keparation and vortex formation 37 Separation is also important for the lifting properties of an aerofoil. At small incidence angles (up to about 10°) the flow does not separate on either side and closely approximates frictionless conditions. The pressure distribution for such a case (“sound” flow, Mig. 2.191) was given in Wig. 1.14. With ineroasing incidones there is danger of separation on the suction side of tho acrofoil, because the pressure in crease becomes steeper. Fora given angle of incidence, which is about 15°, separation fivally occurs. ‘The separation point is located fairly closely beliind the leading edge. The wake, Fig. 2.19b, shows a large “dead-water” area. The frictionless, lift-ereating flow pattern lias become disturbed, and tle drag lias become very large. ‘Tlie be- ginning of separation nearly coincides with the occurretice of maximum lift of the aorofoil. Strnetural aerodynamics. Flow around land-based bluff bodies, such as struc- tures and buildings, is considerably more complex than flow around streamlined bodies and aircraft. The principal cause of complication is the presence of the ground and the shear created in the turbulent wind as a consequence. The interaction between the incident shear flow and the structure produces coexisting static and dynamic loads [8, 9, 10]. The Muctuating forces produced by vortex formation and shedding can indice oscillations in the strictures at their natural frequencics. ‘The flow patterns observed on a detached rectangular building is shown schem: cally in Fig. 2.20. In front of the building there appears a bound vortex which aris from the interaction of the boundary layer im the sheared flow (dV/dz > 0) and the ground, ‘There is, furthermore, strong vortex shedding from tle sharp corners of the building and a complex wake is created behind it. So far no theoretical methods have been developed to cope with this extremely complicated flow pattern, It is, therefore, necessary to resort to wind-tunnel studies with the aid of adequately scaled models. Fig. 2.19a.b. Flow around an acrofoil, aftor Prandtt-Tictjons[27]. a) ‘sound’ flow, h) flow with separation: 38 IT. Outline of boundary-layer theory: Fig. 2.208 Fig. 2.20, Overall view of flow pattern (schematic) aroand a rectangular struc: ture [34]. a) Side view with foreward bound vortex. in the stagnation zone and a separated roof boundary layers b) upwind face and vortex shedding from the tho windward corner of the Ywake roof SN evorter tine S Fig. 2.21, Aerofoil and. cir Gircutar cylinder cular cylinder drawn in such relation to each other as to produce the sane drag in parallel flows (parallel to axis of symmetry of aerofoil) of the same velocity, Acro- foil; Laminar aerofoilNACA 63¢—021 with laminar SperofoisWACA 63,~02t boundary layer. Drag coef- | ficient cng = 0-006 at Rr 108 to 10%, Fig. 17.14. Circular cylinder: Drag coefficient cp = 1-0 at Re = 104 to 105; Fig. 1.4. Thus the ratio of the chord of the acrofoil, 1, to the diameter, d, of the cylinder is Yd = 1-0/0-006 = 167 we MWe 11674 Wig. 2.220 Fig. 2.22. The Reynolds dye experiment. Flow in water made visible by the injection of a dye, after W. Dubs [12]; a) laininar flow, 50; b) turbulent flow, R= 2520 «. Turl lent flow inn pipe and in a bonudary kayer 39 ‘Ta conclude this section, we wish to discuss a particularly telling example of how effectively it. is possible to reduce the drag of a hody ina stream when the separation of the boundary layer is completely climinated and when, in addition, the body itself is given a shape which is conducive to low resistance. Fig 2.21 illustrates Uke effect of a favorable shape (streamline body) on drag: a ayinmeune aerofoil and a circular eylinder (thin wire) have been drawn here to a relative seale which assures equal drag in streams of equal velocity. ‘The cylinder has a drag coefficient Cy © I with respect to its frontal area (sce also Fig. 4). On the other hand, thedrag coefficient of the acro- foil, referred to its cross-sectianal area, has the very low value of Cp = 0-006. The oxtreniely law drag af the acrofoil is achieved as a result of a carefully chagen prolile which assures that the boundary layer remains laminar over almost the whole of its wetted length (laminar acrafoil). Lv this connexion, Chap. XVII and, especially, lig. 17.14, should be consulted. ¢. Turbulent flow in a pipe and in a boundary layer Measurements show that the type of motion through a circular pipe whieh calculated in Scction Td, and in which the volocity distribution was parabolic, exists only at low and moderate Reynolds numbers. ‘The fact that in the laminar motion under discussion fluid laminae slide over cach other, and thnt there are na radial velocity components, so that the pressure drop is proportional to the first power of the inean flow velocity, constitutes an essential characteristic of this type of flow. This characteristic of the motion can he made clearly visible hy introducing a dyc into the stream and by discharging it Uhrough a thin tube, Fig. 2.22. At the moderate Reynolds numbers associated with laminar flow the dyo ie visible in the form of a clearly defined thread exterting over the whole length of the pipe, Fig. 2.22a. By inoreasing the flow velocity it. is possible to reach a stage vhen th fluid particles cease to move along straight lines and the regularity of the motion breaks down. ‘The coloured thread becones mixed with the fluid, its sharp outlin becomes blurred and eventually the whole cross-section becomes coloured, Fig. 2.22. On the axial motion there are now superimposed irregular radial fnetuations which effect the mixing. Such a flow pattern is called turbulent. The dye experiment was first varried ont by 0. Reynolds [29], who ascertained that the transition from the laminar to the turhulent type af motion takes place ata definite value af the Reynolds number (c3 al Reynolds munher). The actual value of the critical Reynolds number depends further on the details of the experimental arrangement. in particular on the amount of disturbances suffered by tho fluid before entering, the pipe. With an arrangement. which is as free from disturbances as possible critical Reynolds numbers (7d) exceeding 104 can be attained (iz = denotes the mean velocity averaged ovor the cross-sectional area). With a sharp-cclged entrance the critical Reynolds number becomes approximately (an This value can be regarded as the lower fi below which even strong Rene & 2300 (pipe). (2.8) | far the critical Reynolds number arbances da not cause the flaw ta heeame turbulent. 40 11, Outline of boundary-layer theory In the turbulent region the pressure drop becomes approximately proportional to the square of the mean flow velocity. Tn this ease a considerably larger pressure difference is required in order to pass a fixcd quantity of fluid through the pipe, ag compared with | r flow. This follows from the fact that the phenomenon of turbulent mixing dissipates a large quantily af energy whieh causes the resistance to flay to increase cansiderably, Furthermore, in the ease of turbulent flow the velo- city distribution aver the cross-sectional arca is much more even than in laninar flow. This ciroumstance is also to be explained by turbulent mixing which causes an exchange of momentuin between the layers near the axis of the tube and those near the walls, Most pipe flows which are encountered in engineering appliances occur at such high Reynolds numbers that turbulent motion prevails as a rule. The laws of turbulent motion through pipes will be discussed in detail in Chap, XX. Ina way which is similar to the motion through a pipe, the flow in a boundary layer along a wall also becomes turbulent when the external velocity is sufficiently large. Experimental investigations into the transition from laminar to turbulent flow in the houndary layer were first carried out by J,M. Burgers [6] and B. G. van der Hegge Zijnen [17] as well as by M. Hansen [16}. The transition from laminar to turbulent flow in the boundary layer becomes most clearly discernible hy a sudden and large increase in the boundary-layer thickness and in the shearing stress near the wall, According to eqn. (2.1), with J replaced by the current co- ordinate 2, the dimensionless boundary-layer thickness 4/\/¥ z/Uos becomes constant for laminar flow, and is, as seen from eqn. (2.12), approximately equal to 5. Fig. 2.23 contains a plot of this dimensionless boundary-layer thickness against the Reynolds number U.,2/v. Ab R, > 32 X 105 a very sharp inerease is clearly visible, and 18 16 w Ri 10 Fig. 2.23, Boundary-layer thickness plot tedr against the Reynolds number based onthe current length x along a plate in parallel flow at zero incidence, as mea- sored by Hansen [16] ce. Turbulent flow ina pipe and in a boundary layer 41 an identical phenomenon is observed in a plot of wall shearing stress. ‘The sudden increase in these quantities denotes that the flow has changed from kuninar to turbulent, The Reynolds mmber R, based on the current ed to the Reynolds mmber Ry = U.. d/r based on the boundary Unronych the en Ry 5 YR, as scen from eqn. (21a). Hence to the critical Reynolds number Ton Ry ert = ( Von r= 32-105 (plate) ¥ there corresponds Ry erie & 2800. ‘The boundary layer on a plate is laminar near the leading edge and becontes turbulent further downstream, The abscissa z,,,, of the point. of transition can be determined from the known value of Ry gr Ln the ease of a plate, as in the previonsly discussed pipe flow, the numerical vate of Reva depends to a marked degree on the amount of disturbance in the external flow, and the value R, ge == 3-2 X 105 should be regarded as a lower limit, With exceptionally disturbance-free external flow, values of R, ori, = 10° and higher have been attained, A particularly remarkable phenomenon connceted with the transition from laminar to turbulent flow occurs in the case of blunt bodies, sich as cireular cylinders or spheres, It will be scen from Figs. 1.4 and 1.5 that the drag coefficient of a circular cylinder or a sphere suffers a sudden and considerable decrease neat Reynolds numbers V D/v of about 5 X 105 or 8 x 10% respectively. This fact was first observed on spheres by G. Hiffel [14]. It is a consequence of transition which causes the point of separation to move downstream, bocanse, in the case of a turbulent boundary layer, the accelerating influence of the external flow extends further due to turbulent mixing. Ienee the point of separation which lies ucar the equator for a laminar boundary layer moves over a considerable distance in the downstream direction. In tarn, the dead area deercases considerably, and the preasure distribution beeomes more like that for frictionless motion (Fig, 1.11). The decrease in the dead-water region considerably reduces the pressure drag, and that shows itself as a jump in the curve Cy =/(R). L. Prandtl [26] proved the correctness of the preceding reasoning hy mounting a thin wire ring nt a short distance in front of Lhe equator of a sphore. ‘This canses the boundary layer to become artificially turbulent at a lower Reynolds number and the decrease in the drag coefficient takes place earlier than would otherwise be the case. Figs. 2.24 and 2.25 reproduce photographs of flows which have becn made visible by smoke. They represent the subcritical pattern with a large value of the drag coefficient and the supercritical pattern with a small dead-water arca and a small value of the drag coefficient. The supercritical pattern was achieved with Prandtl’s tripping wire. The preceding oxperiment shows in a convineing manner that the jump in the drag curve of a circular cylinder and sphere can only be interpreted as a boundary-layer phenomenon. Other bodies 1a blunt or rounded stern, (e. g. elliptic cylinders) display a type of relationship between drag coefficient and Reynolds number which is substantially similar. With increasing slenderness the jump in the eurve hecomes progressively less pronounced. For a streamline body, such as that shown in Fig. 1.12 thore is no jump, because no appreciable separation occurs; the very gradual pressure increase on the back 42, TL. Outline of bon lary. layer theory of such bodies ean be overcome by the boundary layer without separation. As we shall also sce later in greater detail, the pressure distribution in the external flow exerts a decisive influence on the position of the transition point. The boundary layer is laminar in the region of pressure decrease, i.e. ronghly from the leading edge to the point of minimum pressure, and becomes turbulent, in most cases, from that point onward throughout the region of pressure increase. In this connexion it is important to state that. separation can only be avoided in regions of increasing pressure when the flow in the boundary layer is turbulent. A laminar boundary layer, in. 2.24. Mow pasta aphere at aan Reynolds numb Fig. 2.25. Flow pasta aphere ato supereri- from Wieselsherger {391 tioat Reynolis numbers from Wieselaberger [39]. ‘The supercritical flow pattern is achic- ved by the mounting of a thin wire ring (tipping wire) as we shall sce later, can support only a very small pressure rise so tJtat. separation would eccur even with very slender bodies. In particular, this remark also applies to the flow past. an aerofoil with a pressure distribution similar to that in Mig. 1.14. In this case acparation is most. likely to occur on the suction side. A smooth flow pattern around an acrofoil, conducive to the creation of lift, is possible only with a turbulent boundary layer, Summing up it may be stated that the sinall drag of slender bodies as well as the lift of acrofoils are made possible through the existence of a turbulent boundary layer. Boundary-layer thickness: Genorally speaking, the thickness of a turbnlent bonndary layer is larger than that of a laminar boundary layer owing to greater energy losses in the former, Near a smooth flat plate at zero incidence the boundary layer increases downstream in proportion to 28 (x = distance from leading edge). It will be shown later in Chap. XXI that the boundary-layer thickness variation in turbulent. flow is wren by the equation 15 I = 037 ("“") 9.37 (R)-15 (2.9) which cor bounda and watr esponds to eqn. (2.2) for laminar flow. ‘Table 2.1 gives values for the yer thickness calenlated from eqn. (2.9) for several typical cases of air flows. c. Turbulent flow in a pipe and in a boundary layer 43 Table 2.1. ‘Thickness of boundary tnyer, 4, at trailing edge of lat plate at zcro incidence in paralicl turbulent flow Voy = free stream velocity: I= length of plate; += kinematle viscosity Ue t [ft/scey | {ft} Air 100 3 == 150 x 10-8 £t/a0e. 200 3 y x se Fs 7 500 25 150 25 Water 5 5 y= 11 x 10-8 ft/sec 10 15 25 150 50 500 Methods for the prevention of separation: nis mostly an undesirable phenomenon because it entails large cnergy losses. For this reason methods have been devised for the artificial prevention of separation, The simplest. method, from the physical point of view, is to move the wall with the stream in order to reduce the velocity difference between them, and hence to remove the cause of boundary-layer formation, but this is very difficult to achieve in engineering practice. Howe: Prandtl} has shown on a rotating cireular cylinder that this method is very offectiv On the side where the wall and stream move in the same direction separation is com- pletely prevented. Moreover, on the side where the wall and stream move in opposite directions, separation is slight so that on the whole it is possible to obtain a good experimental approximation to perfect flow with circulation and a large lift. Another very effective metliod for the prevention of separation is houndiry- layer suction. Jn this method the decelerated tInid particles in the boundary layer are removed through slits in the wall into the interior of the body. With sufficiently strong suction, separation can be prevented. Boundary-layer suction was used on a circular cylinder by L, Prandtl in his first. fundainental investigation into boundary-layer flow. Separation ean be almost completely climinated with suction through a slit at the back of the circular cylinder. Instances of the effect of suction can be seen in Figs. 2.14 and 2.15 on the example of flows through a highly divergent channel. Fig. 2,13 demonstrates that without suction there is strong separation. Fig. 2.14 shows how the flow adleres to the one side on which suc- tion is applied, whereas from Fig. 2.15 it is seen that the flow completely fills the channel cross-section when the suction slits are put into operation on both sides. In the latter case te streainlines assunie a pattern which is very similar to that in frietioules flow. In lator years suction was successfully used in aeroplane wings to increase the lift. Owing to snetion on the upper surface near the trailing edge, the flow adheres J Prandti-Tietjens; Hydro- and Aerodynamics, Vol. UI, Tables 7, 8 and 9. 44 IT, Outline of boundary-layer theory to the aerofoil at considerably larger incidence angles than would otherwise be the case, stalling is delayed, and much larger maximum-lift values are achieved [36]. Aftor having given a short outline of the fundamental physical principles of fluid inotions with very small friction, i.e. of the boundary-layer theory, we shall proceed to develop a rational theory of these phenomena from the eqnations of motion of viscous fluids. The description will be arranged in the following way: We shall begin in Part A by deriving the goncral Navier-Stokes equations from which, in turn, we shall derive Prandtl’s boundary-layer equations with the aid of the simplifications which can be introduced as a consequence of the small values of vis- cosity. This will be followed in Part B by a description of the methods for the integra- tion of these equations for the case of Iaminar flow, In Part C we shall discuss the problem of the origin of turbulent. flow, i. c. we shall d 88 the process of transition froin laminar to turbulent flow, treating it. as a problem in the stability of laminar motion. Finally, Part D will contain the boundary-layer theory for completely developed turbulent motions. Whereas the theory of laminar boundary layers can he treated ag a deductive sequence based on the Navier-Stokes differential equations for viscous fluids, the same is nol, at, present, possible for turbulent flow, because the mechanism of turbulent flow is so complex that it cannot be mastered by purely theoretical methods. For this reason a treatise on turbuleut flow must draw heavil on experimental results and the subject must be presented in the form of a scmi- empirical theory. References. U1) Achenbach, E.: Experiments on the flow past spheres at very high Reynolds numbers, JEM 54, 565-575 (1972), [2] Berger, K., and Wille, R.: Periodic fow phenomena, Annual Review of Fiuid Mech. 4, 313—340 (1972), 13) Borger, I, Bestimmung der hydrodynamtischen Grissen eincr Kérmanschen Wirbelstrasse aus Hitzdrahtmessungen bei kleinen Reynolds-Zahlen, ZPW 12. 41—-59 (1904). [3a) Bearman, P/W.: On the vortex shedding from a circular cylinder in the critical Reynotds number range, JPM 37, 577—585 (1969), [4] Blasius, H.> Gronzschichton in Fliissigkeitcn mit kleiner Reibnng, Diss, Gottingen 1907; Z. Math. u. Phys. 56, 1—37 (1908); Engl. transl. in NACA ‘TM 1250. [5) Blenk, H., Fuchs, D., and Licbers, L.: Ober dic Messung von Wirbelfrequenzen, Luftfuhrt- forschung /2, 3841 (1935). (6) Burgers, J.M.: ‘The motion of a fluid in the houndary layer along a plane smooth surface, Prov. First International Congress for Applied Mechanics, Delft, 113~-128 (1924). ‘eparation of flow. Pergainon Press, Washington D.C., 1970. Application of fluid inechanics to wind engineering --A Freeman Scholar lecture, Trans, ANMI, Fluids Enginecring 97, Ser. 1, 9--38 (1975); sce also: Laborutory simulation of the atmospheric houndary layer, ALAA J. 9, 1746—1754 (1971). Acrodynamies of buildings. Annual Reyiew of Pluid Mech, 8, 75-106 (1970). (8) Cormak, J-E., and Sadch, W.Z,: Wind-tannel siutulation of wind loading on structures, Mceting Preprint 1417, ASCE National Structural Engineering Meeting, Baltimore, Mary- iand, 19-23 April, 1971. [10] Davenport, A.G.; The relationship of wind structure to wind loading. Proo. Conference on Wind Effects on Buildings and Structures, National Physienl Laboratory, ‘Teddington, Middlesex, Cront Britain, 20--28 June 163, Her Majesty's Stationary Office, London, Vol. 1, 54 112 (19665 References 45 U1) Domm, U.: Ein Beitrag zur Stabilitatstheorio der Wirbelstrassen unter Beriicksichtigung endiicher und zeitlich wachsender Wirbelkerndurchmesser, Ing.-Arch. 22, 400 — 410 (1954). 2) Dubs Wasi a1 Dasa W.: Uber den Fiuflusa laminarer und turbutenter Stromung auf das Rantgenbild you und Nitrobenzol. Helv. phys. Acta 12, 169 -- 228 (1939). WAV, and Karlsson, 8.1 F.: On the phenomenon of vortox street breakdown. 48, GOT -627 (1971). Mel, (2.; Sur la résistance dos aphéres das Pair on mouvement, Comptes Reudus £55, 1597 (1912), [14a] Pérsching, W.AW.: Acroclastische Probleme an Hochbaukenstruktionen in freier Wind umstrémung, Vulkan-Vorlag, Essen, Haus der Technik, Part 347, 3--18 (1976). U5) Primberger, R.: Experimentellc Untersuchungen an der KérmAnachen Wirhelstrasse, ZW 5, 355 — 359 (1957). [16] Hanson, M.: Die Goschwindigkeitsverteilung in der Grenaschicht an der lingsangestriniten ebctien Platte. ZAMM 8, 185—199 (1928); NACA ‘TM 585 (1930). (17) van der Hegge Zijnen, B. jeasurcihents of the velocity distribution in the boundary layer along plane surface, Thesis Dolft 1924, {18} Heinemann, H,J.. Lawaczeck, O,, and Biitefisch, K.A.; Kérmin vortices and their fre- quency determination in the wakes of profiles in tte sub- and transonic regiine, Symposinin ‘Trausonicnn 11 Gdttingon, Sept, 1975. Springer Verlag, 1976, pp. 75 ~82;8ec also: AGAIRD- Conference Proc, No. 177. Unsteady Phenomena in Tarbomachinery (1975). (19) Jlucho, W.H.; Hinflugs der Vorderwagenform auf Widerstand, Giertmoment 0 kraft von Kastenwagen, ZFW 20, 341351 (1972). (20) von Kérmén, Th.; Uber don Mechanisms des Widerstandes, den cin bowegter Korper in einer Fhissigkcit orzcugt. Nachr, Gos, Wiss. Gottingen, Math. Phys. Klasse 509-517 (1911) and 647-~556 (1912); sce also Coll, Worka 1, 324—338, {21) von Kérmén, Th,, and Rubach, H.: Ober den Mechanisnuis des Flissigkeits. und 1 atandes. Phys. Z. 13, 49-59 (1912); sco also Coll. Works 7, 3 (22) Lin, C.C.; On periodically oscillating wakes in the Oxcon appr . Mises Anni- versary Voluine, Studies it Mathematics and Mechattica, Academic Proxs, New York, 1950, 170~17 4) ftwider- Luftwiderstandemessungen am Votkswagen-Licferwagen, Automobiltechnische (1951). -f Mroulal number of bodies and their aystems (ju Russian). Strojnicky Casopis 9 1975), 124) Griffin, 0. M., aud Raniberg, S. E.: ‘The vortex-stroct wakes of vibrating cylinders, JIM 66, 553 —576 (1974). [25) Prandtl, L: Ober Miissigkeitsbewegung bei sehr kleitter Reibung, Prov. 3rd Intern, Math, Congr. Heidelberg 1904, 484-491, Reprinted ins Vier Ablundlungen zur Mydrodynamik und Acrodynamik, Géttingen, 19275 sec also Coll, Works 17, 575-84; Bugl, transt, NACA "TM 452 (1928), [26] Prandtt, 1.2 Der Luftwiderstand von Kugeli. Nachr. Ges, Wiss, Gétlingen, Math, Phys. Kinase, 1914. 177 1903 sce also Coll, Works 11, 597 - GOR, [27] Prandtl, L., aud Tietjens, O.: Hydro- wad Acroweehanik (based on Proudtl’s lectures). Vol. J and iJ, Bertin, 1920 and 1931; Engl, transl, by 1., Rosenttead (Vol, /) and J.P, den Hartog (Vol. 1), New York, 1934, [28] Rolf, E.F., and Sintmous, L-F.G.: ‘The frequencies of exldlics generated by the motion of cirenlar cylinders through a fluid, ARC RM 917, London (1924). [29} Reynolds, 0.: An experimental itwvestigation of the circumstances which determine whether the motion of water shall be direct or sinuons, and of the law of resistance in parallel channels, (hil, Trans, Roy, Soc. 174, 982 (1883); sce also Seientifie Papees 2. 51. [30] Ribner, H1, 8., Etkins, B., and Necly, K. K.: Noise research in Canada; Physical and bio- acoustic. Proc, First Int. Congress Aero, Sei, Madrid, Pergamon Press, London, Vol. 7, 393441 (1959), [31] Roshko, A.: Experiments on the flow past « civeular eyti JIM 10, 345—356 (1961). {32} Roshko; A.: Ou the developmicut of turbulent wakes from vortex streets, NACA Rep. TIT (1954). der at very high Reynolds wumber, 46 I], Ontline of boundary-layer theory {32a} Rosenhead, L.,; The formation of vortices from a surface of discontinuity. Proc. Roy. Soe. A 134, 170 (1931), . [33] Rubach, H.: Ober die Untstehung und Forthowegung des Wirbelpaares bei zytindrischen KOrpern. is ttingen 1914: VDI-Forschiungsheft 185 (1916), [33a] Sarpkaya, ‘T.: Au inviscid model of two.dintensional vortex shedding for transient. aud asymptotically steady flow over an inclined plate. JFM 68, 100-128 (1975). (34) Sadieh, W.Z., and Cermak, J ‘Torbutenee effect ou wall pressure (netuations. J. Eng. Mech, Div, ASCH 98, No. BMG, Proc. Paper M148, 18) 18 (1872). [35] Schlichting. 11; Acrodynamiache Untersuchungen anv Kraftfaltrzcugen. Rep. Techn, Hoch. achule Bramnschweig, 1 (1954), [35] Schronk, O.: Vermuche mit, Abraugelliigeln. Laftfahetforschmg X77, 10 27 (1935). 137] Strouhal, V.; Ober cine besondere Art der ‘Tonerregung, Aun, Phys. and Cheutie, New Sovios 5, 216-251 (1878), [38] Timme, A,: Ober die G hwindigkeitsverteitung i Wirbeln, Ing. Arch. 25, 206 —225 (1957). [88] Wedemeyer. E.: Ausbildnng cinos Wirbelpaares an den Kanton einer Platte. Ing.-Arch, 30, 187 200 (1961), (30) Wiesctaberger, C.: Dor Luftwiderstand von Kugoln, ZPM 5, 140--144 (1914). CHAPTER TIT Derivation of the equations of motion of a compressible viscous fluid (Navier-Stokes equations) } to fh flow a. Fundamental equations of motion and continuity app Wo shall now proceed to derive the equations of motion of a compressible, viscous, Newtonian fluid. In the general case of three-dimensional motion, the flow field is specified by the velocity vector win jodkw where 1, », 1 are the three orthogonal components, by the pressure p, aid by the density Q, all conceived as functions of the coordinates 2, y, z, and time t, For the determination of these five quantities there exist five equations: the continuity equation (conservation of mass), the three equations of motion (conservation of momentum) and the thermodynamic equation of state p = f(9).? The equation of continuity expresses the fact that for a unit volume there is a balance between the masses entering and Ieaving per unit time, and the change in density, [1 the case of non-steady flow of a compressible fluid this condition leads to the equation: Pe ob diva = 2 4-div(gu) = (3.1) whoreas for an incompressible fluid, with g = const, the equation of continuity assumes the simplified form 0. (3.1a) divw The symbol Dg/Dt denotes here the substantive derivative which consists of the local contribution (in non steady flow) Jg/a, and the convective contribution (due to translation), nmgrad 9. Sixth Edition this chapter has been revised by the Translator at the Author's invitation, ‘tional variable, a further equation is ¢ form of the First Law of Thermo- + Inthe + Ef the equation of state contains temperature as an ad supplied by the principle of the conservation of energy in ti dynamics; ef, Chap, XT, 48 IIL. Derivation of the equations of motion of a compressible viscous fluid The equations of motion are derived fron: Newton’s Second Law, which states that the product of mass and acceleration is equal to the suin of the external forces acting on the body. In fluid motion it is necessary to consider the following two classes of forces: forces acting throughout: the mass of the body (gravitational forces) d forces acting on the boundary (pressure and friction). If F = @ g denotes the gravitational foree per unit volume (¢ = vector of acceleration duc to gravity) and P denotes the force on the boundary per unit volume, then the equations of inotion can be written in the following veetor form oe (3.2) with EN 4G 44S body foree (3.3) and P=iP,t iP, +k Py — surface foree . (3.4) syntbol Dee/Dt denotes here the substantive acceleration which, like the substan- live derivative of density, consists of the local contribution (in non-steady flow) @w/ét, aud the convective contribution (due to translation) die/dé = (axgrad) wt Dw aur a dw Dt ~ a di dt ‘The body forces are to be regarded as given external forees, but the surface forces depend on the rate at which the fluid is strained by the velocity field present in it. ‘Phe eystem of forces delormines a state of stress, and it is now our task to indicate the relationship between stress and rate of strain, noting that it can only be given empirically. In our present derivation we shall restrict. attention to isotropic, Newtonian flnids for which it may be assumed that: this relation is a linear one. All gases and many liquids of interest. in boundary-layer theory, in particular water, belong to this class. A fluid is said to be isotropic when the relation betwoen the components of stress and those of the rate of strain is the same in all directions; it is said to be Newtonian when this relation ia linear, that is when the fluid obeys Stokes’s law of friction, In the case of isotropic, elastic solid bodies, experiment. teaches that. the state of stress depends on the inagnitude of strain itself, most engineering materials, obeying Hooke’s linear law which is somewhat aualogous-to Stokes’s law. Whereas the relation between stress and strain for an isotropic clastic solid involves two con- stants which characterize the properties of a given material (c. g. clastic mochulus and Poisson’s ratio), the relation between stress and rate of strain in an isotropic fluid involves a single constant (the viscosity, jz) as long as relaxation phenomena do not it, as we shall sce in See. Ife. ' } In order to express the vector (us grad) «ln an arhitrary system of coordinates, the following general relation shonkT be used (wrpead) we ~ grad $e? — a x curler, where w? rw, b. General stress system in a deformable body 49 b. General stress system in a deformable body In order to writo down expressions for the surface forces acting on the boundary, ct us imagino a small pnrallepiped of volume dV = da dy dz isolntex| instantaneously from the body of the fluid, Fig. 3.1, and let its lower left-hand vertex coincide with the point x, 7,2. On the two faces of area dy+dz whieh are perpendicular to the x-axis there aot two resultant stresses (vectors = surface force per unit area): gy pz and pz-+ £5 de respectively . (3.5) Fig. 3.1. Derivation of whe expressions for the stress tensor of an inhomogencons 5 stress aystoin and of ils aymmetry in the absence of a volumetric distribution of * local moments (Subscript x denotes that the stress vector acts on an clementary plane which is perpendicular to the x-direction.) Similar terms arc obtained for the faces dx «dz and dz « dy which are perpendicular to the y- and z-axes respectively. Henec the three net components of the surface force are: a plane | direction 2: 57% -da- dy: dz p, oY. . . nom OW gy de dy de ap, no» 2 gh da dy: dz and the resultant surface force P per unit volume is, therefore, given hy 2, Py, om a tay toa (3.6) The quantities py, py, pz are vectors whieh can be resolved into components por- pendicular to each face, i.c., into normal stresses denoted by o with a suitable subseript indieating the direction, and into components parallel to cach face, i. ¢. into shearing stresses denoted by t. ‘The symbol for a shearing stress will he provided 50 IIT, Derivation of the equations of motion of a compressible viscous fluid with two subsoripts: the first subscript indicates the axis to which the face is per- pendicular, and the second indicates the direction to which the shearing stress is parallel, With this notation we have Paid, +f tay +k tre Py = ity: 1 fty +k tye (3.7) Pe = ite bj ty thor. ‘The stress system is scen to require nine sealar quantities for its description, Theso nine quantities form a stress tensor. The set of nine components of the stress tonsor is sometimes called the stress matrix: Se try Tre Waly. of tw). (3.8) Ue Ty ‘Tho stress tensor and the corresponding matrix are symmetric, which means that two shearing stresses with subscripts which differ only in their order are equal. This can be demonstrated with reference to the equations of motion of an element of fluid, Tn general, its motion can be separated into an instantaneous translation and an instantaneous rotation, and ouly the latter needs to be considered for our purpose. Denoting the instantaneous angular acceleration of the element by «(c,, y, @:), we can write for the rotation about, the y-axis that toy Uy = (ty, dy de) dr — (ty, dar dy) dz = (typ — Tx) dV where d/, is the elementary moment of inertia about: the y-axis. Now the moment of inertia, dJ, is proportional to the fifth power of the linear dimonsions of the pa- rallelepiped, whereas its volume, dV, is proportional to their third power. On contract- ing the clement. to a point, we notice that the left-hand side of the preceding equation vanishes faster than the right-hand side, Hence, ultimately, 1, 0 ry ~ Tyr if @, is not to become infinitely large. Analogous equations ean be written for the remaining two axes, and the symmetry of the stress tensor ean thus be demonstrated. Tt. is clear from the argument that the stress tensor would cease to be symmetric if the {uid developed a local moment. which was proportional to its volume, dV. The latter may ocour, for example, in an electrostatic field. Owing to the fact that, Ty = Tye ae teri Tye Tey (3.9) the stress matrix (3.8) contains only six different!stress components and hecomes symmetrical with respect to the principal diagonal ; Os Tay Tre N=(ty oo tw). (3.10) Tes ty O b. General stress aystom in a deforinable body 51 The surface force per unit volume ean be calenlated from eqns. (3.6), (3.7), and (3.10) and becomes pail(Z 24. oe 42 ) pees comp. x Or, ao, oT bog ) tates comp. 4 (3.10a) ér, a, p (See ve Oe) 0 +4(% +5 4 *) comp. 2. face face fave ye mm ory Introdneing the expression (3.10a) into the equation of motion (3.2), and resolving into components we have: oma x a (Gr + Oe 4 =) De emare(? rey 4. Fu 4 a) G11) Dw Ore, , a, an, oman (Mea te, ). If the fluid is “frictionless” all shearing stresses vanish; only the normal stresses remain in the equation, and they are, moreover, equal. Their negative is defined as the pressure at the point x, y,z in the fluid: ty =te=tye= 0 Oe = Oy =O, = —p. In such a hydrostatic stress system, the fluid pressure is equal to the arithmetien! mean of the normal stresses taken with a negative sign. Sinec measurements which lead to the establishment of the thermodynamic equation of state are performed under such conditions, the fluid being at rest, this pressure is identical with the thermodynamic pressure in the eqaation of state. It is convenient to mtroduce the arithmetical mean of the three normal stresses — their sum being called the trace of the stress tensor — as a usclil mumorieal quantitiy in tho easo of a viscous fluid in a state of motion also. It is still called the pressure, but its relation to the thermodynamie pressure requires further investigation. Although it then ceases to be equal to a particular stress whieh is normal to the surface, it has the property of being invariant with respect to transformations of the system of coordinates, as it is an invariant of the stress tensor, heing defined ag doz + oy + 02) =—p. (3.12) We shall sec in Sce. Ife that it remains equal to the thermodynamic pressure in the absenec of relaxation. 52 II. Derivation of the equations of motion of a compressible viscous fluid ‘The system of the three equations (3.11) contains the six stresses dg, dy, de, Ty Tre tye ‘The next task is to determine the relation between them and. the strains so as to enable ns to introduce the velocity components x, v, w into eqn. (3.11). Before giving this relation in See, ITld we shall investigate the system of strains in greater detail. ined in flow When a continous body of fluid is made to flow, every clement in it is, goncrally xpeaking, displaced to a new position in the course of time. Daring this motion clements of fluid heeome strained, and since the motion of the fluid is completely determined when the velocity vector w is given as a function of time and position, w = w(z,4,2,4), there oxist kinematic relations between the components of the rate of strain and this function, The rate at which an clement of fluid is strained depends on the relative motion of two points within it. We, therefore, consider the two neighbouring points A and B which are shown in Fig. 3.2. Owing to the presence of the velocity field, point A will be displaced to A’ in time dé by a distance s = w dt; nee, however, the velocity at B, imagined at a distance dr from A, is different, point. B will move to B‘ displaced from B by s -|- ds = (ww + div) dt, More explicitly, if the components of velocity have the values u,v, wat A, then, at the neighbouring point B, the velocity components will be given to first order by the Taylor-scries expansions, au ufdusn 1. a t a ov ybedy =o 3. vebdy =o St tz (3.13) ew wf dues 10 bg da +} ative motion of point. B with respect to A is described by the following matrix of nine partial derivatives of the local velocity field au ou Oe an by be a a av : ac dy ae (3.13) ow aw ew de by ae Vig. 3.2. Relative displacement ¢, The rate at which a fluid element is strained in flow 53 It is convenient to rearrange the expressions for the relative velocity components du, dv, dw from equ. (3.13) to the form dus (€, dx -b &gy dy + &,, 2) + (dz — Edy) (ly = (Ey, da 4- €, dy -b &y, dz) 1 (fda — & dz) (3.14) day = (Bz, dat sb bay dy Le €, dz) 4 (Edy — yd), | it being casy to verify that the new symbols have the following meanings el eu . La. du). a faw ) eu\ < be by bre Be ; (2 | ms +( 4 ®); » fe . 1 (aw 1 (aw a bye] bye by bye 3 (3 + 3 (" F s) ; Sel: 1 [au aw rr xy be ola: + ie (3.154) and aw 1 (au aw - au It is noted that the inatrix é, is symmetric, so that 3 by = Eat by = Eyes (3.160) and that £, 7, are related to the components of the vector w = curlw (3.15) Fach of the new terms can bo given a kinematic interpretation, and we now proceed to obtain it. Since wo concentrate our attention on the immediate neighbourhood of point A, and since interest is centred on the motion of B relative to A, we shall place point. A at the origin, and interpret dx, dy, dz as the coordinates of point Bin a Cartesian system of coordinates. In this maunor, the expressions in oqns. (3.14) will define a ficld of relative velocities in which the components du, do, diy are lincar frnctions of the space coordinates. In order to understand the meauing of the differeut terms in the matrix (3.16a) and in eqna. (3.15b), we proceed to interpret them one by one. ‘The diagram in Fig. 3.3 represents the field of relative velocities when all terms except @u/ax vanish on the assumption that @u/dx > 0. ‘The relative velocity of any point B with respect to A is now Ou du = (2) ax, and the field consists of planes 2 = const which displace themselves uniformly with a velocity which is proportional to the distance dz away from the plane x An elementary parallelepiped with A aud B at its vertices placed in auch a velocity fiold will be distorted in extension, its face BC receding from AD with an increasing vation of the equations of inotion of a compressible viscous fnid Fig. 3.3. Local distortion of fluid clement when dujac > 0 with all other terme being equal to zero; uniform extension in the x direction velocity. Thus é, represents the rate of elongation in the x-direction suffered by the clement. Similarly, the additive terms é, = dv/dy and é, = Aw/Az deacribe the rate of clongation in the y- and z-direetions, respectively. It is now casy to visualize the distortion imparted to a fluid clement: by the simullancons netion of all three diagonal clements of matrices (3.18) or (3.152). ‘Phe clement expands in all three directions, and the change in the length of its Unree sides produces n change in volume at a relative rate a a» 20 {az + Maz ah fw, hy a fae +e ay} = deay de da dy az di au 4 a) OY _ div, (3.16) ae + ay | a é= to first order in the derivatives. During this distortion, however, the shape of the clement, described by the angles at its vertices, remains unchanged, since all right angles coutinue to be that way. Thus é describes the local, instantancons volumetric dilatation of a (luid element. When the fluid is incompressible, é = 0, as must be expected. In a compressible fluid the continuity equation (3.1) shows that ay 1 De : divw =~ 9 ape G17) that is that the volumetric dilitation, the relative change in volumé, is equal to the negative of the relative rate of change in the local density. + The relative velocity field presents a different appearance when one of the off-diagonal terms of matrix (3.13a), for example du/dy, has a non-vanishing, say positive, valuc, The corresponding fickd, sketched in Fig. 3.4, is one of pure shear strain, A rectangular clement of fluid centred on A now distorts into a parallelogram as indicated in the diagram. The original right angle at A changes at a rate measured by the angle yxy = ((0u/dy) dy dt}/dy, that is at a rate @ujdy. When both dujoy e. The rate at which a fluid element is strained in flow a Bi 7 fr Fig. 3.4. Local distortion of fluid clentent —- when au/ay > 0 with all other terms being equal to zero; uniform shear defornmtion, yh 4 Yarat Fig. 3.5. Local distortion of Haid olentent, when bay = Eye = } (Oulu) + (Gn/ex)} > 0 with all other terms being equal to zero; distortion im shape. (The diagrain has been drawn for dujéy = dujax ) Fig. 3.6. Local distortion of finid element when = 4 {(@uféx) — (Oufo)) + 05 instantancous rigid-body rotation 56 IIL. Derivation of the equations of motion of a compressible viscous fluid and anjéx have positive nonvanishing values, the right angle at A will distort owing to the superposition of two motions, the state of affairs being illustrated in Fig. 3.5. It is clear that. the right. angle at A now ports at twice the rate desetibed by two of the off-diagonal torins of matrix (3.162), In general, the three off-diagonal orms &,, = b yz, ge = Esns Ud Ez, = éy, describe the rate of distortion of a right angle located in # plane normal to the axis the index of which does not appear as a subscript. ‘The distortion is volume-preserving and affects only the shape of the element. Civenmstances are again different in the particular caso when au/dy = — dv/ax illnatrated in Vi Fron: the preceding considerations and from the fact that. now é,, = 0 we can “infer at onee that the right angle at A remains undistorted. “This is a'also clear from the diagram which shows that the fluid clement rotates with respect. to the referenes point A. Instantaneously, this rotation occurs without distortion and can he described as a rigid-body rotation. The instantancons angular velocity of this rotation is (Gefaxydedt _ a ded ~ ve ay IL. is now easy to sec that the component. ¢ of 4 curl w from eqn. (3. the vorticity of the velocity field, represents the angular velocity of thi rigid-body rotation, and that 5b), known as instantancous eu ey Jo. oe an or) fur = 8 lag |} be, as describing the rate of distortion in shape, the term cal (a _ = aloe ~ &y describing the rate at which the clement of flnid participates in a rigid-body rotation, ‘The linearity of eqns. (3.13) or of the entirely equivalent equs. (3.14) signifies that the most general case arises by a superposition of the simple cases just described. ‘Therefore, if aliention is fixed on two neighbouring points A and B in a body of fluid which snstains a continuons velocity ficld w(2,4,z), the motion of an element of fluid surrounding these two points can’ be uniquely decomposed into four compo- nent. motions : c. The rate at which a flnid element is strained in flow 57 (a) A pure translation described by the velocity components u,v, of w. (b) A rigid-body rotation described by the components & 7,0 of 4 curl w. (c) A volumetric dilatation described by ¢ = div w, the linear the dircetion of the axes being described hy é,,é, and é, () A distortion in shape described by the components é,, ete with mixed indie Only the last. two motions prodnee an intrinsic deformation of a (nid element. surrounding the reference point A, the first two causing a mere, general, displacerhent of its location. ‘The elements of matrix (3.15) constitute the components of a symme' tensor known as the rate-of-strain tensor; ita mathematical properties are analogous to those of the equally symmetric stress tensor. It is known from the theory of elasticity [3, 7] or from general considerations of tensor algebra [11] that with every symmetric tensor it is possible to associate three mutually orthogonal privcipat axes which «etermine three mutually orthogonal principal planes that is a privileged Cartesian syste of coordinates, Ii this systein of coordinates, the stress vector or the instantancons motion in any one of the principal planes is normal to it, that is, parallel to one of the axes. When such a special system of coordinates is used, the matrices (3.10) or (3.15a) retain their diagonal terns only. Denoting the values of the respective components by symbols with bars, we would be dealing with the matrices o, 0 0 % 0 0 o % 0 and [05 4 0 (3.18) 0 0 & 0 0 & It should, finally, be remembered that such n translormation of coordinates does not affect. the sum of the diagonal terms, so that g, | 0, 16,6, +6, | G5 (3.198) and by + be Ley be, (=e =divw), (B.19b) because they conxtitnte invariants of the tensors, as already intiniated earliel Viewed in such two aystems of coordinates (both denoted by bars). an clement Vig. 3.7. Principial axes for stress and rate of strain 58 JIL, Derivation of the equations of motion of a compressible viscous fluid of flnid is stressed in three mutnally perpendicular directions, and its faces are displaced instantaneously also in three mutually perpendicular directions, as suggested by Figs. 3.7a and b, ‘This docs not, of course, mean that there exist no shearing stresaes in other planes or that. the shape of the element remains undiatorted. d. Relation between stress and rate of deformation Tt should, perltaps, be stressed once more that the equations whiek relate the surface forces to the flow ficld must be obtained by a perceptive interpretation of experimental ilts and that our interest is restricted to isotropic and Newtonian fluids, ‘The considerations of the preceding section provided us with a useful mathe- matical framework which allows us now to state the requirements suggested by experiments in a somewhat more precise form, When the fluid is at rest, it develops a uniform field of hydrostatic stress (negative pressure -— p) which is identical with the thermodynamic pressure, When the fluid is in motion, the equation of state still determines a pressure at. every point (“principle of local state” [4]), and it ig convenient to consider the deviatorie normal streases go, = 6, | Py oy yh Ds Gf =o, DG (3.20) together with the imehanged shearing stresses. The six quantities so obtained constitute a symmetric stress tensor the existence of which is due to the motion beeause at rest. all its compononts vanish identically, From what has been sail before it follows that the components of this deviatoric tensor are created solely by the components of the rate-of-strain tensor, that is to the exclusion of the com- ponents 1, », 2 of velocity as well as of the compononts , 1, ¢ of vorticity. This ia equivalent to saying that the istantancons translation [component motion (a)} as well as the instantaneous rigid-body rotation (component. motion (b)} of an element of flnid produce no surface forces on it in addition to the existing com- ponents of hydrostatic pressure, ‘The preceding statement, evidently, merely repre- souls a precise local formulation of what. we expect to observe when a finite body of fluid performs a general motion which is indistinguishable from that of an equivalent rigid body. We thus conclude that the expressions for the components GO, Oy'. 2. +s Tex OF the deviatoric stress Lensor can contain in them only the velocity gradients @u/dx, ..., Qw/Pz in appropriate combinations which we now proceed to determine, ‘These relations are postulated to be linear; they must remain unchanged by a rotation of the system of coordinates or by an interchange of axes to ensure isotropy. lsotropy also requires that. at every point in the continuum, the principal axes of the stress tensor must coincide with the principal axes of the rate-of-strain tensor, for, otherwise, a preferred direction would be introduced, ‘The sintplest way to achieve our aim is to scleet an arbitrary point in the continuum and to imagine that the local system of coordinates #, 7,2 has been provisionally so chosen as to coincide with the three common principal axes of the two tensors, The com- ponents of the velocity field in this system of coordinates are denoted by %, 3, m. It is now clear that isotropy can be secured only if cach one of the three normal is mule to depend on the component of rate of strain the direction ntresses 7/0 5.02 d. Relation between stress and rate of deformation 59 of whicl: conicides with it and on the sunt of the three, each with a different factor of proportionality: Thus we record, directly in terms of the apace «lerivatives, that, G21) ‘The quantities u,v, w and €, 7, ¢ do not appear in these expressions for the reasons just explained. In each expression, the last term represents the appropriate rate of linear dilatation, that is, in essence, a change in shape, and the first term repre- sents the volnmctrie dilatation, that is the rate of change in volume, in essence, a change in density. The factors 2 jn the last terms are not essential, being merely convenient to facilitate the interpretation, as we shall see later. The factors of proportionality, 2 and A, two in all, must be the same in cach of the three preceeding equations to secure isotropy. It is easy to see that an interchange between any two axes, that is an interchange of any of the three pairs of quant (BF), (1,2) leaves the set of relations invariant, as they must be in an isotropic medinm. Moreover, the preceding is the only combination of spatial gradients which possesacs the required propertics. If the reader cannot sec this directly, he may consulta more rigorous proof in a treatise on tensor algebra (or c.g. [Il] p. 89). ‘Tho relations in eqns. (3.21) car be re-written to apply in an arbitrary system of coordinates by performing a general rotation with the aid of the appropriate linear transformation formulac, We shall refrain from putting down the detailed steps because, thongh tedious if performed directly, they are quite straightforward, They become simple if tensor ealculus is used. ‘The approriate direct formulae may be found in refs. [8, 6, 7], whereas their tensorial counterparts are given in ref. [11]. Such a derivation would show that og =Adivw $24 % y= Adivw +2 p (3.224) =Adivw 42h an ou cute a(S +) oe ) (3.220) 44 wy az} 60 TEE, Derivation of the equations of motion of » compressible viscous fluid where div w has been used for brevity. The reader may notice tho regularity with which the indices 2, y,2, the components 2, ,w, and the coordinates x, y,2 are permuted f. Applying these equations to the simple ease represented in Fig. 1.1, we recover eqn. (1.2) and so confirm that the preceding more general relation reduces to Newton's law of frietion in the case of simple shear and does, therefore, constitute its proper generalization. At the same time, we identify the factor » with the viscosity of the fluid, amply discussed in Sec 1b, and, incidentally, justify the factor 2 previously inserted into eqns. (3.21). The physical significance of the second factor, A, requires further disenssion, but we note that. it plays no part. in an ineompressible fluid when div w= 0; it then disappears from the cquations altogether, and so is seen to be important for compressible fluids only. e. Stokes’s hypothesis Although the problem that we are about to discuss has ariset more than a century and a half ago, the physical interpretation of the second factor, A, in eqns, (3.21) or (3.22a,b) and for flows in whicl: div w docs not vanish identically, is still being disputed, even though the value which should be given to it in the working equations is not. This numerical value ia determined with the aid of a hypo- thesis advanced by G, G. Stokes in 1845 [13]. Without, for the moment, concerning ourselves with the physical reasons which justify Stokes's hypothesis, we first state that according to it, it is necessary to assume BA 2n=0, o Aa 2p. (3.23) This relates the value of the factor A to the viscosity, j, of the compressible fluid and reduces the number of propertics which characterize the ficld of stresses in a flowing compressible fluid from two to one, that ia to the same number as is required for an incompressible fluid. Substituting this value into equa, (3.22a), we obtain the normal components of deviatoric stress: , 2 . Ou oy = — Ff adivi top oy 5 edivw + 2p 5 (3.24) of =~ Fp adivi tone, ' + The aboyc act of six equations can be contracted to a single one in Cartesian-tensor notation (with Einstein's snmmation couvention): , Ove (2 avy Oy ~ Ady diy + ye Or, + ie where the Kronecker delta 64) = 0 for i+ jand dyy — | fori = j. ). (i,j, & = 4, 2,3) £. Bulk viscosity and thermodynainie pressure 61 the shearing stresses remaining unchanged. Making use of eqns. (3.20), we obtain the so-called constitutive equation for an isotropic, Newtonian fluid p= pf pdivw 1 2p m 2 dive 42p 2 (3.25) 3 Ff # oy 2 . eur . fdiv 4 2p Ty = (3.25b) tye = Ter = in its final form, noting that p represents the local thermodynamic pressuret. Regarded as a pure hypothesis, or even guess, eqn. (3.23) oan certainly be accepted on the ground that the working equations which result from the substitution of eqns. (3.25a,b) into (3.11) have been subjected to an unusually large mimber of experimental verifications, even under quite extreme conditions, as the reader will concede after having studicd this book, Thus, even if it should not represent, the state of affairs exactly, it certainly constitutes an excellent approximation. Since the deviatorie components are the only ones which arise in motion, they represent those components of stress which produce dissipation in an isothermal flow, Utore being further dissipation in a temperature field duc to thermal conciction, Chap, XL, Furthermore, since the factor 2 occurs only in the normal components 6,/.0,',0;' Which also contain the thermodynamic pressure, equs. (3.20), it becomes clear that. the physical significance of A is connected with the mechanism of dissi- pation when the volnne of the fluid clement is changed at a finite rate as well aa with the relation between the total stress tensor and thermodynamic pressnre, £. Bulk viscosity and thermodynamic pressure We now revert to the general discussion, without necessarily accepting the validily of Stokes's hypothesis, but. confine it lo the case when no shearing stresses are involved, because their physical significance and origin is clear. Consequently, “ In the compact tensorial notation we would write ao | avy Poy + n( + by 7 vx. Og= *) GAR =12 9). 62 IIL, Derivation of the equations of motion of a compressible viscous fluid we consider a flnid system, say the aphere shown in Fig. 3.82 which is subjected toa uniform: normal stress, @, on its boundary, Tn the absence of motion @ is obviously equal and opposite in sign lo the thermodynamie pressure, p. ‘Taking the sum of the three equations (3-21) and utilizing eqs. (3.20), we find that F- pia piv, (3.26) and notice that our equations reflect this fact, as alrendy pointed out earlier. Now, the qneation poses itself as to what this relation should be in a general flow field. y Fig. 3.8, Quasistatic compression and oscillatory motion of a spherical mass of fluid When the system is conipressed quasistatically and reversibly, wo again recover the previous case because ten div w —» 0 asymptotically, We note that in such cases the rate at which work is performed in a thermodynamically reversible process per unit, volume beoomes W=pdivw (3.26) which is the same as dV Wop Ge (3.26b) in the notation enstomary in thermodynamies, Whon div w is finite, and the firid is compressed, expanded or made to oscillate,’ ata finite rate, equality between @ and — p porsists only if the coefficient. whi gn (3.27) vanishes identically (Stokes’s hypothesia); otherwise it docs not, If yz’ -|0, the oscillatory motion of a spherical system, Fig, $8, would produce dissipation, even if the temperature remained constant throughout the bulk of the gas, The sane would be truc in the case of expansion or compression at a finite rate, For thia reason, the coefficient yu’ is called the bulk viscosity of the fluid: it represents that. property, like the shear viscosity jz for deformation in shape, which is responsible for cnergy dissipation in a fluid of uniform temperature during a change in volume £. Bulk viscosity and Ukermodynantic pressure 63 at a finite rate, The bulk viscosity would thus constitute a accom property of a compressible, isotropic, Newtonian luid needed to determine its coustil itive e« and would have to be measured ia addition ta ye. Lt is evident that ation w= 0 tnplicg po a 1-10 implies p | ; Thus the acceptance of Stokes's hypothesis is equivalent to the assmmyption that the thermodynamic pressure p is oual the one-third of the invariant sum of normal strosaes even in cases when compression or expansion proceeds at a tinite nite, Furthermore, il is.also equivalent. to the assumption that the oscillatory motion of a large spherical eystem would be reversible if it were isothermal, More detailed considerations in terma of the concepts of thermodynamics as it, applies to irrever- sible processes in continuous systems can be found in the works of J. Meixner [8], 1. Prigogine [12] aud 8. R. de Groot. and P, Mazur [I]. In order to determine under what. conditions the bulk viscosity ofa compressible (nicl vanishes, it is necessary to have reeonrse to experiment or to the methods of statistical thermodyuamies which permit. us to calculate transport coefficients frou fire principles, The direct. measurement of bulk viscosity is very difficult to per. form, and no definitive results are in existence, Statistical methods for deuse gases or liquids have not yet been developed to a point which world allaw ns to make a complete statement on the subject. It appears, however, that the bulk viscosity vanishes idontically in gases of low density, that is under conditions when only binary collisions of molecules necd to be taken into account, In dense gases, the numerical value of bulk viscosity appears to be very small, ‘This nieans that: equs. (3.26a,b) continue to describe the work in a continuous system in the absence of shear to an excellent degree of approximation and that dissipation at constant. tenipera- ture, even in the general case, occurs only through the intervention of the devia- lorie strosses, ‘Thus, once again, we are led to Stokes’s hypothesis and so to eqn. (3.26). This conclusion docs not. extend to thrids which are capable of undergoing relaxation processes by virtne of a local departure from a state of chemical equili- hrinm [1,8]. Such relaxation processes oceur, for example, when a chemical reaction can take place, or, in gases of complex structure, when a comparatively slow transfor of energy betaveen the translational and rotational degreca of freodom on tie one hand, aud the vibrational degrees of freedom ou the other, becomes possible. ‘Tins when relaxation processes are possible, the thermodynamic pressure is no longer equal to one-third of the trace of the atress tensor, [t is sometimes argued that the adoption of Stokes's hypothesis, that ix the supposition that the bulk viscosity of Newtonian flnid vanishes, does not, accord with our intuitive fecliug that, a aphere of fluid whose boundary oscillates so hat there is a cyclic sequence of compression and expansion, Fig. 3.8b, would dissipate no energy. This would, indeed, be the ease, as is casily seen from: the preceding argument, because the dissipative part of the stress fickl vanishes wnder sich con- ditions. It must, however, not be forgotten that such a conclusion is valid only if the temperature of the sphere of gas were to be kept constant, during the oscillation throughout the whole volume, Normally this is impossible, Consequently, an oscillating sphere of gas will soon develop a temperature field and cnergy will be dissipated down the existing temperature gradients (5. 64 IEE, Derivation of the equations of motion of a compressiblo viscous fluid g. The Navier-Stokes equations With the aid of equs, (3.20) the nou-viscoua pressure terms can be separated iy the equation of motion (3.11) so that they become Du _ ap, (202. ety wn) em ax (SE + ay + a Dv _ ap ey Oey vw) Cin =~ Yay t ( az + ay + a (3.28) Do _ ep Oe OTF ee +(32 + 5 Introducing the constitutive relation from eqns. (3.24) we obtain the resultant, surface force in terms of the velocity components, ¢. g. for the a-direction we obtain with the aid of eqn. (3.10a): aa,’ er, er, Orey 1 Ofer a ty te =? Tan op é eu 2 . a eu, év é ew, ou Pam — ae toe ont 5 paivw| +f [a(#+2) | ta [ (32 +2) | and corresporling expressions for the y- and z-componenta, In the gencral case of a compressible flow, the viscosity j must be regarded as dependent on the space coordinates, because varies considerably with temperature (Tables 1.2 anc 12,1), and the changes in velocity and pressure together with the heat due to friction bring about considerable temperature variations, The temperature dependence of viscosity (T) must. be obtaincd from experiments (cf. See. XTITa). If these expressions are introduced into the fundamental equations (3.11), we obtain Du é Bu 2 é eu ev é ew eu ein = as [1(2% gd w)| +8 [n(% +2) + [(3e + 2) Dv gay 24. a aw @| (au , w ein we — 5 aru) +e [ol& +R) + lo +2) ev ee eRanedloot fare} g[oles)] es[(-4a)] (3.298, b, et These very well known differential equations forin the basis of the whole science of fluid mechanics, They are usually referred td as the Navier-Stokes equations, + In indicial notation: ( am) Cla tM az, ep, af fam , ay 2 aol ~ Kem 5 elle + Be 3 80 Set} (ik = 2.8). g. The Navier-Stokes equations 65 Itis necessary to include here the equation of continuity which, as seen from equ, (3.1), assumes the following form for compressible flow: 2 , Pou) , aor) , alow) at be + ay ! a (3.30) ‘The above equations do not give a complete description of the motion of a com- pressible fluid because changes in pressure and density effect temperature variations, and principles of thermodynamics must, therefore, once more enter into the con- siderations, From thermodynamics we obtain, in the first place, the characteristic equation (equation of state) which combines pressure, density, and temperature, aud whieh for a perfect gas has the form p—eRT 31) with R denoting the gas constant and 7' denoting the absolute temperature, Secondly, if the process is not isothermal, it is further necessary to make use of the energy equation which draws up a balance between heat and mechanical energy (Iiral Law of Thermodynamics), and which furnishes a differential equation for the temperature distribution. The energy equation will be discussed in greater detail in Chap. XI The final equation of the system is given by the empirical viscosity law y2(7), its dependence on pressure being, normally, neglected. In all, if the forees X, ¥,Z are considered given, there are sever equations for the seven variables x, », 1, p, @, 7, fe For isothermal processes these reduee to five oquations (3.29a,b,c), (3.30) and (3.31) for the five unknowns 1%, », &, p, 2. Incompressible flow: ‘The above aystem of equations becomes further simplified in the case of incompressible fluids (9 = const) even if the temperature is not constant, First, as already shown in eqn. (3a), we have div w = 0, Secondly, since temperature variations are, generally speaking, small in this case, the viscosity may be taken to be constant. The equation of state as well as the energy equation become superfluous as far as the calculation of the field of flow is concerned, The field of flow oan now be considered independently from the equations of thermodynamics, The equations of motion (3.29a,b,c) and (3.30) ean be simplified and, if the acceleration terms are written out fully, they assume the following form: eu eu bu eu o(% tugte a 4 wit) Ou 5a bya avy, av gat og + (3.32a, b,c) Ow , tw Se 4 Oe | Me) (3.33) } This condition ia more nearly satisticd in gases than in Fiquida. 66 ILL. Derivation of the equations of motion of a compreasibie viscous fluid With known body forces there are four equations for the four unknowns 4, ¥, 10, p. If vector notation is used the simplified Navier-Stokes equations for incompres- sible flow, eqns. (3.32a,b,c), can be shortened to Dw oy = Ferd pt pyre, (3.34) where the symbol 7? denotes the Laplace operator, 72 = 02/dx? -- 02/ay? -+ 02/022. The above Navier-Stokes equations differ from Euler’s equations of motion by the viscous terms ye \/? w. ‘The solutions of the above equations become fully determined physically when tte boundary and initial conditions arc specified. In the case of viscous fluids the condition of 10 slip on solid boundaries must be satisfied, i. e., on a wall both the normal and tangential components of the velocity must. vanish: % =O, % = 0 onsolid walls. (3.35) The equations under discussion were first derived by M. Navier [9] in 1827 and by §. D. Poisson [10] in 1831, on the basis of an argument which involved the considoration of intermolecular forces. Later the same equations were derived without the use of any such hypotheses by B. de Saint Venant [14] in 1843 and by G. G. Stokes [13] in 1845. Their derivations were based on the same assumption as made here, namely that the normal and shearing stresses are linear functions of the rate of deformation, in conformity with the older law of friction, due to Newton, and that. the thermodynamic pressure is equal to one-third of the sum of the normal stresses taken with an opposite sign. Since the hypothesis of linearity is evidently completely arbitrary, it is not a priori certain that the Navicr-Stokes equations give a true description of the motion of a fluid. It is, therefore, necessary to verify them, and that can only be achieved by experiment. In thia connexion it should, in any ease, be noted that the enormous mathematical difficulties encountered when solving the Navier-Stokes equations have so far prevented us from obtaining a single analyticsolution in which the convective ternts interact in a general way with the friction terms. However, known solutions, such as laminar flow through a circular pipe, as well as boundary- I flows, to be discussed later, agree so well with experiment that the general validity of the Navier-Stokes equations can hardly be doubted. Cylindrical coordinates: We shall now transform the Navier-Stokes equations to cylindrical coordinates for future reference. If r,¢, z denote the radial, azimuthal, and axial coordinates, respectively, of a threc-dimensional systom of coordinates, and 2, 2%, 0, denote the velocity components in the respective directions, then the transformation of variables [3, 11] for the case of incompressible fluid flow, eqns. (3.33) and (3.34), leads to the following system of equations: 2», ay, uy a, Ug? ae, 2 2 8 i o(5 ey - ? fot Ee ap oy 1 ph ee 2 Ny" Oy, 7 r or rt Th st Bgt 8 BP (3.368) g. The Navier-Stokes equations 67 av, ayy | My Ary oY, ou, 6 tH Me $ olgh tu gt the + y t)= _ 1 ap ary 1% %y 1 Fey 2 ey, a) =Fs— pat a (Ber a at as apt bs ag + as) (3900) ee, ay, % , er, ote ty a 1, 1 Oy ) . =F soa tT agit (3.366) a, | % 1 buy oy, atst+ymetaH? (3.36d) The stress components assume the form 0, a 1%, meant wank eH HER 1%, on 1 &, oo —p tan (he 4 re =n (Ge 42 (3.37) eu, a, av, %=—P+2yzs re =n (Ze + S). Curvilinear coordinates: Tt is often useful to employ a curvilinear rystem of coordinates which is adapted to the shape of the body, In the ease of two-dimensional flow along a curved wall, it is possible to select a coordinate system whose abscissa, z, is measured along the wall, the ordinate, y, being measured at right angles to it, Fig. 3.9. Thus the curvilinear net consists of curves which are parallel to the wall Fig. 3.9. Two-dimensional boundary layor along a curved wall and of straight lines perpendicular to them. ‘The corresponding velocity components are denoted hy 1 and », respectively. The radius of curvature at position « is denoted by R(x); it is positive for walls which are convex outwards, and negative when the wall is concave. The appropriate form of the complete Navier-Stokes equations has been derived by W. Tollmien [15]. They are: 68 Derivation of the equations of motion of a compressible viscous fluid Ou R my vu R_ 3 a a by ate ag RY vi: + Oty Ow I ou u + lta ryt ot bat TRG ay Tee F388) - 2k be R aR Ry AR du (Ri pt de (Ray de bi +o de Hy}: ev 4 R u av oh av we i op 4 {3 2R au at Rey ar ay Rey oe dy or (Rk + yt oe 1 av Re ary ” He yay | Tae oF UE a oe yah Ay Ry AR aw (Ro4 ys dr (Ro+ y)* cee ey R bu " Tw =0. (3.380) he stress components are Ro Ou v oom alery oe beta) y= — pe 2p ee (3.39) ty iy (2 | M\ay hy | and the vorticity [see equ. (4.5)} becomes Tey I R eu ony (FG : (3.40) References UJ de Groot, S.R., and Masur, P.: Nonequilibrium thermodynamics. North-Iotland Publ. . 1062. [2] Foppl. Ac: Vorlesungen jiber technische M 13] Hopf, Le: Ziihe Milistigkeiten. Contribution and K. Scheel, ed.), Bertin, 1927. [4] Kestin, J.2 A course in thermodynamies. Vol. 7, Blaipdell, 1966. {5] Kestin, Je: Etude thermodynamique des phénoménes irréversibles. Rop. No. 66--7, Lab. hanik. Vol. 5, Teubner. Leipzig, 1922. Handbuch der Physik, Vol. VT (H. Geiger a Adrothermiqne, Meudon, 1968. [6] Lamb. Hf: Hydrodynamics. 6th ed., Cambridge, 1957; also Dover, 1945. {7} Love, A. th: the mathematical theory of clasticity. 4th od., Cambridge Univ. Press, ton2. {8} Meixner, +b. sad Rei, 1G. Pherniodynamik der irroversiblen Prozesse. Contribution to Maudbuch der Physik, Vole J/1/2 (8. Mligge, ef.), Springer, 1959, pp. 413-523. References 69 [9] Navier, M.: Mémoire sur tes lois duu mouvemont des finides. Mém. de Acud. de Sci. 6, 389--418 (1827). [10} Poisson, S.D.: Mémoire sur tes équationa générales de Péquilibre et du monvement des corps solide i = dol Heole polytochn. 73, 139 (86 (1831). a , W. 8 of continua, Ginn & Co., 1961. [12] Prigogine, {.: Elude thermodynamique des phi 8 irrévorsiblos, Danod- Desuer, 1947. u On the theories of internal friction of fluids in motion. ‘ ps. Camber, Soc. &, 287—305 (18! [4] de St. Venant, B.: Note & joindre un mémoire sur la dynamique des fides. Comptes Rendus 17, 1240—1244 (1843) U5] ‘Tollmien, W.: Grenzachichttheorie. Handbuch der Bxper.-Physite, Vol. JV, Part, 1, 241-287 + (1931). CHAPTER IV General properties of the Navier-Stokes equations Before passing on to the integration of the Navier-Stokes equations in the following chapters, it now seems pertinent to «liscuss some of their gencral properties. In doing so we shall restrict ourselves to incompressible viscous {uidds. a. Derivation of Reynolds's principle of si Navier-Stokes equati Until the present day no general analytic methods have becoine available for the integration of the Navier-Stokes equations. Furthermare, solutions which are valid for all valnes of viscosity are known only for some particular cases, c.g. for Poi- seuille flow through 2 circular pipe, or for Couctte flow between two parallel walls, one of which is at rest, the other moving along its own plane with a constant velocity (sco Fig. 1.1). For this reason the problem of calculating the motion of a viscous fluid was attacked by first tackling limiting cases, that is, by solving pro- bleins for very large viscosities, on the one hand, and for very small viscosities on the other, becanse in this manner the mathematical problem is considerably sim pli- fied. However, the case of moderate viscosities cannot be imterpolated between these two extremes. Even the limiting eases of very large and very small viscosities present great mathematieal diffieulties so that researel into viscous ftuid motion proeeededd to a large extent by experiment. In this connexion the Navier-Stokes equations furnish very uscful hints which point to a eonsidcrablo reduetion in the quantity of experimental work required. It is often possible to earry out experiments on models, which means that in the experimontal arrangement a geometrically similar model of the actual body, but redueed in scale, is investigated in a wind tunnel, or other suitable arrangemont. This always raises the question of the dynamic similarity of fluid motions which is, evidently, intimately connected with the ques- tion of how far results obtained with models can be utilized for the prediction of the behaviour of the full-scale body. 1 As already explained in Chap. I, two fluid motions are dynamically similar if, with geometrically sinilar bonndlaries, the velocity fills aro geometrioally similar, i. ¢.,if they have geometrically similar stroamlines. This quostion was answored in Chap. T for the ease in whieh only inertia and viscous forces take part in the process. Tt was found there that for the two motions a. Derivation of Reynolds's principle of similarity 7 the Reynolds unmibers mast. be equal (Reynolds's principle of similarity). ‘This conclusion was drawn by estimating tho forees in the stream; we now propose to. deduce it again directly from the Navier-Stokes equations, ‘The Navier-Stokes equations expross the condition af equilibrium, namely that. for cach particle there is equilibrinm betweon body forces (weight), surface forces and inertia. forecs, The surfaco forces cousist of pressure forces (normal forces) and friction forces (shear forces). Bor ¢ important only in cases when there is a free surface or when tho density distribution is inhomogencons. In th case of a homogenecons Muid in the absence of a free surface there is cquilibrinn botweon the weight of cach purticle and its hydrostatic hnoyancy forco, in the same way as at rest. Hence in tho motion of a homogencous fluid, in the abscuec of a surface, body forees cau be cancelled if pressure is taken to mean the differ hetween that in motion and at rest. In the following argament we shall restrict our attention to cases for which this assumption is true becanse they are the most. in- portant ones in applications, ‘Thus the Navier-Stokes equations will now contain only forces due to pressure, viscosity, and inertia. Uniler these assumptions and conventions the Navier-Stakes equations for ant incompressible fnid, restricted to steady flow and in vectar fori, sinrplity to 2 (uxgrad) w = — grad p +p y2w. (4.1)t This differential equation must he independent of the chaice of the units for the various physical quantities, such as velocity, pressure, ete., which appear in it. We now consider flows about two geometrically similar hodies of different linear dimensions in streams of different velocities, c. g., flows past two spheres in which densities and viscosities may also be different. We shall investigate the con- dition for dynamic similarity with the aid of the Navier-Stokes equations. Kvi- dently, dynamic similarity will prevail if with a suitable choice of the units of length, tine, and force, the Navier-Stokes eqn. (4.1) is so transformed that. it hecanies identical for the two flows with geometrically similar boundaries. Now, it is possible to free oneself from the fortuitonsly selected units if dimensionless «quantitios arc introduced into eqn. (4.1). This is achieved by selecting ocrtain snitahle charac teristic magnitudes in the flow as our units, and by referring all others to them, Thus c. g., the free-stream velocity and the diameter of the sphere catr he selected as the respective nits of velocity and length. Let V,1, and p, denote these characteristic reference magnitudes. If we now introduce into the Navier-Stokes equ. (4.1) the dimensionless ratios velocity AW’ = longths = X = ; , P ressure P=? , pressure Pr t See footnote on p. 48. 72 Ww. Joneral propertics of the Navier-Stokes equations we obtain ay Vey 1 lw e), (egead) = — TF grad P+ Ne PW or, dividing by 9 V3/t: (Fe grad) Ws — gui arad Pl fe (4.2)} The fluid motions mder consideration can become ly the soluti aed in terms of the respcetive dimensionles ariables are identical, This requires that for hott motions the respective dimensionless Navier-Stokes equations differ only by a factor common to all terms, The quantity p,/g V2 represents the ratio of pressure to the donble of the dynamic head and is unimportant for the similarity of the 4vo motions hecause in incompressible flow a change in pressure causes no change in vohnne. The second factor g V liye is, however, very important and mnst assume the same value for both motions if they are to he dynamically similar. Hence dynamic similarity is assured if for the two motions Mil ovals Ay Pe This principle was discovered hy Osborne Reynolds when he investigated fluid motion through pipes and is, therefore, known as the Reynolds principle of similarity. ‘The dimeusionless ratio: evi Vlg h , (4.3) called the Reynolds mimber. Here the ratio of the dynamic viscosity p, to the density g, denoted by » = p/o, is the kinematic viscosity of the fluid, introduced carlicr, Summing up we can state that flows about geometrically similar bodies are dynamically similar when the Reynolds numbers for the flows are equal, ‘Thos Reynolds's similarity principle has been deduced onee more, this time from the Navier-Stokes equations, haviug heen previously derived first from an estimation of forces and secondly from dinensional analysis. b. Frictionless flow as “solutions” of the Navier-Stokes equations It may be worth noting, parenthetically, that the solutions for incompressible frictionless flows may also be rogarded as exact. solutions of the Navier-Stokes equations, bocanse in such cases tho frictional terms vanish identically. In the case of incompressible, frictionless flows the velocity vootar ean he represented ax the gradient of a potential: w= grad, whore the potential @ satisfies the Laplace oqnation vid=0. We then also have grad (V? 6) — V2 (grad @) = 0, that is, V?w = 0. { Soe footnote on p. 48. e. The Navier-Stokes equations interpreted as vorti Inport equations 73 ty t Thus the frictional terms in eqn. (4.1) vanish identically for potential flows, but generally speaking both houndary conditions (3.35) for tho volocity cannot then bo satisfied simultanconsly. Th'the normal coniponent must nesnme preseribed valucs along bonndacy, then, in potential flow, the tngential component is Uhoroby detorminod so thit the no slip condition ennnat he satisficd at tho same time. For this reason one cannot regard potontinl flows as physically meaningful solutions of the Navior-Stokos equations, because they do not antisfy Ut prescribed boundary conditions. ‘There exist, however, an important exception lo the procetiing state- ment which occurs when the solid wall is in motion and when this condition does uot, appl The sinplest particular case is that of flow past a rotating cylinder when the potential sol tion does constitute ineaningful golution to the Navier-Stokes equations, as explained in greater detail on p. 80. The reailor may rofer to two papers. one by G. Hamel [4] and one hy. J. Ackeret (1], for further details. ‘The following sections will be restricted to the consideration of plane (two-dimensional) flows heeause for stich cares only is it possible to indicate some general properties of Une Navier- Stokes equations, and, on the other hand, plane flows constitute by fur the largest class of probleins of practical importance, ec. The Navier-Stokes equations interpreted as vorticity transport equations In the case of two-dimensional non-steady flow in the x, y-plano tho velocity vector becomes w= 6u(r,yt) + fo(x,y,t). and the system of equations (3.32) and (3.33) transforms into au au) atu ate (2s + en) ov a | a Be ( + oe (44a, b, ¢) which furnishes three equations for u, v, and p. We now introduce the vector of vorticity, curl w, which reduces to the one component about the z-axis for two-dimensional flow: 1 _ 1 fav aw x Letwaomont (@—). ay Frictionless motions are irrotational so that curl w = 0 in such eases. Eli- minating pressure from equs. (4.4a, b) we obtain aw eo ew aw aw et ae +P ay =” (i ay (4-6) or, in shorthand form Dey vio. (47) be This equation is referred to as the vorticity transport, or transfer, equation, 1t states that the substantive variation of vorticity, which consists of the local and convective 74 IV. General properties of the Navier-Stokes equations terms, is equal to the rate of dissipation of vorticity through friction, Eqn, (4.6), together with the equation of continuity (4.40), form a system of two equations for the two velocity components u and », Finally, it is possible to transform these two equations with two unknowns into one equation with one unknown by introducing the stream function y(z. y). Putting va (4.8) we sce that the continuity equation is satisfied automatically, In addition the vorticity from eqn. (4.5) becomes o=—4 Ppt (4.9) and the vortioi y transport equation (4.8) becomes oy dviy ap aviy + Fe ON EAE ty. (4.10) = = Fig. 4.1 Patterns of motion 1a viscous flow past, a sphero at different Reynolds numbers R-- H'Djw derived from the vorticiky transport. equmtion (4.10) by V. G, Jonson [5]. a,b,c, Patterns of streamlines; dl. 0, £, Distribntion of borti PV = const ad no separation bee separation at 6 = 171° of separation at 6 = 148° f ‘This equation is sometimes called the Poisson eqpintion for tho atreunfunction y. ©. The Navier-Stokes equations interpreted as vorticity transport equations 15 In this form the vorticity transport equation contains only one unknown, y. The left-hand side of eqn. (4.10) contains, as was the case with the Navi equations, the inertia terms, whercas the right-hand side contains the fri terms, It is a fourth-order partial differential equation in the stream fret lis sohition in general terms is, again, very difficult, owing to its being non-linear. V. G, Jenaon [5] found a sohition to the vorticity transport. equation (4.10) for the case of a sphere by numcrical integration. The resulting patterns of strcan- Hines for different Reynolds muimbers are seen plotted in Fig. 4.1 which also contains diagrams of the distribution of vorticity in the flow field. The smallest. Reynolds number inchided, R = 5 in Figs. 4.la and 4.1d, corresponds to the case when the viscous forces by far outweigh the inertia forces and the resulting flow can be de- scribed as creeping motion, See. TVd and Chapter VI. In this case the whole flow field js rotational and the patterns of streamlines forward and aft are nearly identical, As the Reynolds nmnber is inercased the sphere develops on ita rear a separated region with back-flow and the intensity of vorticity is progressively more concentrated near the downatream portion of the sphere, whereas in the forward portion the flow be- comes nearly irrotational. The flow patterns under consideration which have been dedveed from the Navier-Stokes equation, allow vs to recognize the characteristic changes which take place in the stream as the Reynolds number is made to inerease, even if at the highest Reynolds number reached, R = 40 in Figs. 4.1¢ and 4.1f, the boundary layer pattern has not yet had a chance to develop fully. Vory extensive experimental investigations of the wake behind a circular cylinder in the range of Reynolds numbers 5 < R < 40 are described in two papers by M. Cou- tanceau and R. Bouard [lo, 1d} who covered both stoady and unsteady flows. ‘The development of very efficient clectronic computers in modern times has made it possible to solve the Navier-Stokes equations for flow past. geometrically simple bodies by purely numerical methods. In order to do this, the differential eqvations are replaced by difference equations. The numerical techniques used for this parpose will be explained in Sec. [X1. Without disenssing this matter here in any depth, we quote one interesting result. Figure 4.2 shows the flow past a rectangular plato placed at right angles to the stream calculated by J.B. Fromin and F. H. Harlow [3]. At the back of the plate there forms a vortex strect similar to that behind a circular cylinder shown in Figs. 1.6 and 2.7. Figure 4.2ashows an experimentally determined pattern of streamlines, whereas Fig. 4.2b represents the calculated field, both for a Reynolds number Vd/v = 6000. The agreement between the two patterns ia remark. ably good, in apite of the fact that in this range of Reynolds mmbers the flow aequires an oscillatory character, Fig. 1.6. The earliest attempts to obtain avelt mimerical solutions to the Navier-Stokes equations can be traced to A, Thor [6[ who per- formed such calculations for a circular cylinder at the low Reynolds munbers R ~ 10 to 20. Later, the calculations were carried to R = 100 [2]. As the Reynolds mumber increases, the degree of difficulty of such numerical integrations inereases steeply. In this connextion it is worth consulting the comprehensive summary by A. ‘Thom and C. J. Apelt [7}, as well as the work of C. J. Apelt [La] and DN. de G. Allen and R. V. Southwell [Ib] and of If, B, Keller and If, Takami [6al. 76 IV. General properties of the Navier-Stokes equations Fig. 4.20, Fig. 4.2. Pattern of streamlines hehind a rectangular flat plate (If/d = 146) placed. ht augle to the flow at a Reynolds number R= V fl/y = 6000, after JK, Fromm and FH. Harlow (3). (i hoight of plate, d = thickness of plat: a) streamline pattern determined experi- mentally, b) streamline pattern calenlated by nit merical integration of the Navier-Stokes equation for T = t F/I = 2-78 (1 = time from atart of motion), Numerteal integra- tion performed on an IBM 7090 computer 4.2b a. The limiting case of very large viscosity (very small Reynolds number) In very slow motions or in wotions with very large viscosity the viscons forces are consilerably greater than the inertia forces because the latter are of the order of the velocity sqnared, whereas the former are lincar with velocity. ‘To a first approximation it. is possible to neglect the inertia terms with respect to the viscous terms so that from eqn. (4.10) we obtain Vips= " (4.11) ‘This is, now, a linear equation which is considerably more wanenable to mathe- inatical treatment than the complete equ. (4.10). Flows desoribed by equ. (4.11) proceed with very sinall velocities and are sometimes called creeping motions, The ‘The limiting ease of very small viacous forces 7 omission of the inertia torus is pomnissible from the mathematical point of view because the order of the eqvation is not. Uereby reduced, so that with the simpli- fied differential equ. (4.11) it ix possible to satisfy as many boondary conditions as with the full equ. (4.10). Creeping motions can also be vegarded as solotious of the Navier-Stokes equations in the Liamiting case of very small Reynolds numbers (Ro ~ 0), he the Reynolds number represents the ratio of inertia to friction forces. Solutions of oqn. (4.11) for the creeping motion of a viscous fluid were foul by G. G. Stokes in the case of a sphere and by EH. Lamb in the case of a circnlat cylinder. Stokes’s solution can be applicd to the falling of particles af mist. in air, or to the motion of small spheres in a very viscous oil, when the velocities are so small that inertia forees can be neglected with good accuracy, Fnrthermore, the hyplvolynamic, theory of lubrication, ic. the theory of the motion of lubricating oil in the very narrow channel between the journal and bearing uses thie simplified equation of motion as its starting point. In the latter ease it will he observed that if Ube velocities are not very small, the very small clearance heights, and the rela- ‘osity of the oil, enspre that the viscous forees are much larger than the inerbin Torees. However, apart Irom tho theory of lubrication, the Beld of apph- cation of Uhe theory of ereeping motion is fairly insignificant, AUIS e. The miting case of very small viscous forces (very large Reynolds numbers) From the point of view of practical applications the second extreme case, namely that. of very sinall viscons forees in eqn. (4.10) compared with the inertia forces, is of far greater importance. Since the two most important flnids, namely water and air, have very small viscosities, the case under consideration occnrs, gencrally speaking, already at moderately high velocities. ‘This is the limiting case of nery large Reynolds numbers (R > 00), In this case the process of mathematical simplification of the differential eqn. (4.10) requires a considerable amount, of care. It is not. permissible simply to omit the viscous terms, i. ¢., the right-hand side of eqn. (4.10). This would reduce the order of the equation from four to two, and the solution of the simplified equation could not be made to satisfy the full boundary conditions of the original eqnation, The problem which was ovtbined in the preceding sentences belongs essentially to the realm of boundary-layer theory. We now propose to discuss briefly the general statements which can be made about the solutions of the Navier-Stokes equations for the special case of small viscous forces as com- pared with the inertia forces, that is in the limiting case of very large Reynolds nombers, : ‘The following analogy may serve to illustrate the character of the soluti of the Navier-Stokes cquations for the limiting case of very small viscosity, i. c., of very small friction terms, as compared with the inertia terms. ‘The temperature distribution 0(x, y) about a hot body in a fluid stream is described by the follow- ing differential equation, Chap. XII; ne 78 TV. General properties of Wig Navier-Stokes ecpiationa 20 20 20 e70 ao oc (Fe poe) we (EY. (4.12) Here 9, ¢, and & denote the density, specific heat, and conductivity of the Mnid ‘ely; @ is the difference between the local temperature and (hal at a very ie temperature, 7, is constant, and equal to i ne(x, 9) ad (ae, y) in eq. (4.12) is aasstaned ion on the boundaries of the body defined mplest case it. ie constant with respect to point. for an clementary volume. ‘The left- by 7p = Teo is prescribed and in the space and Lime but, generally speaking, it varics with both. l'rom the pltysic of view equ. (4.12) representa the heat balan hand side represents Ule quantity of heat, exchanged by convection, whereas the right-hand side is the quantity of heat exchanged by conduction. ‘The frictional heat generated in the fluid ia neglected. If 7'q > T.. Wie problem is that. of determining the temperatare field around a hot body which is cooled. By inspection it is seen that eqn. (4.12) is of the same form as eqn. (4.6) for the vorticity w. In fact they become identical if the vorticity is replaced by the temperature differenee and the kinematic viscosity » by the ratio k/ ¢ known as the thermal diffusivity. The boundary condition 0 = 0 at a large distanee from the body corresponds to the condition =0 for the undisturbed parallel stream also at a large «listance from the body. Uenee we may expect tat the solutions of the two equations, i. of vorticity and that of temperature around the body will be similar in character. Now, the tem perature distribution around the body may be perceived intuitively, to a certain extent. In the limiting ease of zero velocity (fluid al reat) Ue influ the heated body will extend uniformly on all sides. With very small veloci fluid around the body will atill be affected by it in all directions. With increasing velocity of flow, however, it is clearly scen that the region affected by the higher lemperature of the body shrinks more and more into a narrow zone in the imme- diate vieinity of the body, and into a tail of heated fluid behind it, Fig. 4.3. Fig. 4.3. Analogy between temperature and vorticity distribution in the neigh- bourhood of a body placed in a stream of thud 1), b) Limits of regton of Increased Lemperntare 3) for matt velocities 3) for intge velocities of flow ‘The solntion of eqn. (4.12) must, as mentionéd, be of a character similar to that for vorticity. AU sniall velocities (viscous forecs large compared with inertia forces) there ty in Ute whole region of flow around the body. On the other hand for large velocities (viscous forces mall compared with inettia forces), we may expect. a field of flow in whieh vorticity is confined to a amall layer along the surface of the hody and in a wake beltind the body, whereas the rest of the field of flow e. The limiting ease of very small viscous forces 79 remains, practically speaking, free from vorticity (see Fig. 4.1). It is, therefore, to be expected that in the limiting case of very amall viscous forces, ic. ab la Reynolds numbers, the solutions of the Navier-Stokes equations are so coustiluted aa to permit. a subdivision of the field of flow into an external region which is fre from vorticity, and a thin layer near the body together with a wake behind it. In the first. region the flow may be expected to satisfy the eqptations of friction flow, the potential flow theory being nsed for ils evalnation, whereas in the second region vorticity is inherent, and, therefore, the Navier-Stokes equations mist be used for its evaluation. Viscous forecs arc important, i.e. of the samo order of magnitude ae inertia forees, only in the second region known as the houndary layer. ‘This concept of a boundary layer was introduced into the science of Amid inechanics by L. Prandtl at the beginning of the present century: it has proved to be very frvitfol. The subdivision of the field of flow into the frictionless external flow and the essentially viscous boundary-layer flow permitted the reduction of the mathe- matical difficnltics inherent. in the Navier-Stokes equations to soch an extent that it became possible to integrate them for a large number of cases. ‘Che description of tbese methods of integration forms the subject of the boundary-layer theory pro- sented in the following chapters. From a numerical analysis of the available solutions of the Navier-Stokes equations it is also possible to show dircctly that. in tho limiting case of very large Reynolds numbers there exists a thin boundary layer in which the influcnee of vis- cosity is concentrated. We shall revert to this topic in Chap. V. The previously discussed limiting case in which viscous forecs heavily outweigh inertia forces (creeping molion, i. c., very small Reynolds number) results in a con- siderable mathematical simplification of the Navier-Stokes equations. By omitting the inertia terms their order is not reduced, but they become linear. The second limiting case, when inertia forees outweigh viscous forces (boundary layer, i. e. very large Reynolds numbers) presents greater mathematical difficulties than creeping motion. For, if we simply substitute » = 0 in the Navier-Stokes equations (3.32), or in the stream-function equation (4.10), we thereby suppress the derivatives of the highest order and with the simpler equation of lower order it is impossible to satisfy simultancously all boundary conditions of the complete differential equations. However, this does not signify that the solutions of such an equation, simplified by the elimination of viscous terms, lose their physical meaning. Moreover, it. is possible to prove that this solution agrees with the complete solution of the full Navier-Stokes equations almost everywhere in the limiting case of very large Reynolis numbers. ‘The exception is confined to a thin layer near the wall — the boundary layer. ‘Thns, the complete solution of the Navier-Stokes equations can be thonght of as consisling of two sohitions, the so-called “outer” solution which is obtained with the aid of Buler’s equations of motion, and a so-called “inner” or boundary-lnyer sohition which is valid only in the thin layer adjacent to the wall. The “inner’’ sohition satisfics the so-called boundary-layer equations which are deduced from the Navier- Stokes equations by coordinate stretching and passage to the limit R -> oo, as will be shown in Chap. VI. The outer and inner solutions must be matched to each other by exploiting the condition that there must exist an overlapping region in which both solutions are valid. 80 IV. General properties of the Navier-Stokes equations f, Mathematical illustration of the process of going to the limit Root Since the preceding argmnent conatihites one of th layer theory, it nay be worth while lo illuatrate the basi chich wan firat given by L. Proidtl*. ce fundamental principles of bomdary- leas by quoting a mathematical analog, Let us consider the damped vibrations of a point-inass described by the differential equation m der +k se Ler 0. Tr | (4.13) Here m denotes the vibrating mass, ¢ the spring constant, & the damping factor, « the length coordinate nteasured from the position of equilibrium, aud f the tine. The initial conditions are agmimed to be =O at r= (4.14) In analogy with the Navier-Stokes equations for the ease when the kinematic viscosity, F, is very small, we consider here the limiting case of very small mass m, because this too catacs the term of the highest order in eqn. (4.13) to become very small. ‘The complete solution of eqn. (4.13) subject to the initial condition (4.14) has the form A {exp (— ¢ tk) ~ exp (~kt/m)}; m > 0, (4.15) x oo 81 In spite of tho simplification, the differential equation (4-20) is one of second degree: it ean be made to satialy the initial condition (4.14) by the choice (4.22) ‘The value of constant Az follows from the suntching to the “outer” sohition, equ. (4.17). ht an overlapping range, that is for moderate values of time, the solitions in eqns. (4.17) and (4.21) must agree, ‘Thus we must have Tin ae (t*) = lim ay (H) co to or, in words: ‘The “outer” limit of nice sohition must be equal Lo the “outer” solution. Condition (4.23) leads at once to Ar=A, (4.24) and 60 bo the inn a(t) = A {1 — oxp(-- ke}. (4.25) ‘The same form can be obtained from the contplete solntion from eqn. (4.15) by expanding the first. Lerm for small values of € and retaiwing the first {erm only, that is by putting lim exp (--et/k) = 1. (1.26) too ‘The two solutions, the outer solution from eqn. (4.17) and the inner solution from eqn. (4.25), together form the complete solution on conclition that cach is used in its proper range of validity. At finite ¢, eqn. (4.15) tends to the outer solution for m — 0. whereas at constant & eqn. (4. tends to the inher solution. ‘The partial solutions give the complete, composite solution which is valid in the entire range of values of { by adding them together, remembering tat the conmon term from eqn (4.23) must. be included only once. that is subtracted from the suin according to the prescription zu (0) = 0 (t) 4 (er) a(t) = roll) + xt) — li im aro(t). (4.27) 9 A graphical representation of the complete solution from eqn. (4.15) is shown in Fig. 4.4 for the ense when A > 0, Curve (a) corresponds to the outer solution (4.17). Curves (h), (¢) and (a) represont solutions of the complete differential equation (412) with mm decreasing from (b) to (d) If we now compare this examplo with the Navier-Stokes equations, we conclude that. the complete equation (4.13) is astalogous to the Navier-Stokes equations for a viscons fluid. whereas the simplitivd equation (4.16), corresponds to Euler's equations for aur ideal tlnid. The Fig. 4.4. Solntions of the vibration equation (4.13), (a) Solution of the simplified equation (4.14), or == 05 (b), (0). (a) represent solutions of the completa differential equation (4.13) wir various values of m. When m is very sinall, solution (a) avprires: boundary-layer eter 82 LV. Goncral propertios of the Navier-Stokes equations n (4.14) plays a part which is similar to the no-slip condition of a real fluid. ‘The latter can bo satislied by the solutions of tho Navior-Stokes equations hut not by those of Enler's equations, ‘The slowly-varying solution ia analogous to the frictionless solution (potential flow) which fails to satisfy the no-slip condition. ‘Tlie fnst-varying solution represents the couttterpart of the boundary-layer solution which is deterinined by the presence of viscosity; it differs from zero only int a narrow zono near tho wall (boundary layor), It is to be noted that the second bonndary condition (no slip at the wall) can only be satisfied if this boundary-layer solution is acldod, thus tnaking the whole solution physically real, This simple example exhibits the same mathematical fenturos as those d receding chapter. It is, tmmely, not permissible singly to omit the vinous in the i tiona when performing Ute process of going over to the limit for very smal viscosity (very large Reynolds number). This ean only be done in Ute integral solution itself. We shall demonstrate lator in greater detail that it is not necessary to retain the full Navier-Stokes equations for the process of finding the limit for Roo. For the sake of mathematical simplification it will prove possible to omit certain torms in it, particularly certain small viscous terms. It is, however, important to note that not all viscous terms can be neglected, as this would depress the order of the Navier- Stokes equations. Roforences [1] Ackeret, J.: Ober exakto Lésungen der Stokes-Navier-Gicichungen inkomprossiblor Missig- keiten bei veriinlerton Gretzbe-lingungen, ZAMP 3, 259-271 (1952), Ua] Apelt. he steady flow of a viscous fluid past a cirenlar cylinder at Reynolds numbers 40 and 44. British ARC RM 3175 (1961). [1b] Allen, D.N. De G., and Southwell, R. V. motion, in two dimensions, of a viseo Math. 8 129.145 (1955). [le] Coutanceau, M., and Bouard, R.: Experimental determination of the main features of the viseous flow in the wake of a circular cylinder in uniforin translation, Part 1. Steady flow. JEM 79, 231—256 (1977). Ud] Coutanceau. M., and Bovard, R.: Experimental determination of the main features of the viseous flow in the wake of a eireular cylindor in uniforin translation. Part 2. Unstoady flow. AFM. 79, 257 ~272 (1977) [2] Dennis, 8. and Zu Chang: Nionerical salutions for ateady flow past a cireular cylinder at Reynolds numbers up to 100, JFM 42, 471 ~-489 (1970), [3] Fromm. J.1, and Harlow, F.1L: Ninncrical solutions of the problem of vortex strect developmont. Phys. of Flnids 6, 975 — 982 (1963); acc. also: ATAA Selected Reprints, Compu- tational Fluid Dynamics (C.K. Chn, cd.), 82--89 (1968) and AGARD Lecture Sories 34 (t974). [4] Uamol, [5] Jenson, V. Soc. London A [5a] Koller, 11.13, and ‘Takami, 1 udies of steady viscous flow abont cylinders, Namerien! solutions of non-li I equations. Proc. Adv. Symp. at Univ. of Wisronsin, Madison. 1966 (1D. Greenspan, ed.), J. Wiley & Sons, New York, 1966, pp. 115-140. [6] ‘Thom, A. 651. 669 (1933), [7] Thom, A., and Apelt. + strand, London, 1961. : Relaxation methods applied to determine the. id past a fixed oylinder. Quart. J. Mech. Appl. Uber die Potentialstrémung ziiher Fliasigkciten. ZAMM 27, 129 139 (1941). Viscous flow round a sphere at low Reynolds numbers (< 40). Proc. Roy. ' low past cireular cylinders at low speeds. Proc. Roy. Soc. London A 141, Ficld computations jn cngineering and physics. Van No- CHAPTER V Exact solutions of the Navier-Stokes equations In general, the problem of finding exact solntions of the Navier-Stokes exua- tions presents insurmountable mathomatical diffioultios. This is, primarily, a con- sequenee of their being non-linear, so that. the application of the principle of super- position, which serves so well in the casc of frictionless potential motions, is excluded. Nevertheless, it is possible to find exact solutions in certain particular casey, when the quadratic convective terms vanish in a natural way. In thia shall devote our attention to the discussion of several exact solutions. Incidentally, it will he shown that in the case of small viscosity many of the exact solutions have a boundary-layer structure which means that the influence of viscosity is confined to a thin layer near the wall. ‘A comprehensive review of solutions of the Navier-Stokes equations has been given by R. Berker [4]. a. Parallel flow Parallel flows constitute » particularly simple class of motions. A flow is called parallel if only one velocity component is different from zero, all fluid particles mov- ing in one dircetion. For example: if the components v and w are zero everywhere, it follows at once from the equation of continuity that Au/ée = 0. which moans that. the component w cannot depend on x. Thus for parallel flow we have w=ulya; v=0; wH0. (5.1) Further, it also follows immediately from the Navicr-Stokes equations (3.32) for the y- and z-direotious} that ap/ay = 0, and dp/az = 0, so that. the prossnre depends only on x. In addition, in the equation for the x-direction all convective terms vanish, Honee au dq, ote atu eget +n (5 +%) (6.2) which is a linear differential equation for w(y, z, !). + In the following argument the term “prosmiro” denotes the difference botweon the total pressitre and the hydrostatic pressure (pressure at rest). ‘This causes the body forces to eaucel, as they are in equilibrium with the hydrostatic pressure. 84 V. Exact solutions of tle Navier-Stokes equations 1. Parallel flow through a straight channel and Couette flow. A very simple solution of equation (5.2) is obtained for the case of steady flow in a channel with two parallel Nat walls, Pig. 5.1, Let the distanee between the walls be denoted by 2 6, so that eqn. (5.2) ean be wri dp de . dz age (5.3) with the boundary condition: 1 -= 0 for y =» 4: b. Since ap/iy = 0 the pressure gra- dient in the direction of flow is constant, as seen from equ. (5.3). Tis dpjda = const and the solution is 1g? (ye — 4), (5.4) US On de ‘The resulting velocity profile, Fig, 5.1, is parabolic. Fig. 5.1, Parallel flow with parabolic velocity distribution Another simple solution of eqn. (5.3) is obtained for the so-called Couette flow between two parallel flat walls, one of whieh is at rest, the other moving in its own plane with a velocity U, Fig. 5.2. With the boundary conditions y=0: u=0; ysh: w=U we obtain the solution _# wary ( _y 5 us Uo ie AE (55) which is shown in Fig. 5.2. In particular for a vanishing pressure gradient. we have us 4 U. (5.58) ‘Thin particular case is known as simple Couette flow, or simple shear flow. ‘The general case of Couette flow is a superposition of this simple case over the flow between two flat walls. ‘The shape of the velocity profile is «letermined by the di- monsionless pressure gradient 1 ne ap Peay (-2) : For P > 0, i.e, for a pressure decreasing in the direetion of motion, the velocity is positive over the whole width of the channel. For negative values of P the velocity a. Parallel flow 85 > Pore arr -04 -02 0 02 04 06 08 10 12 1% s S Fig. 5.2. Conctte flow between two parallel (at walls 17> 0, pressure decrease In direction of wall wl P< 0, presente Increase: 7? = 0, z0r0 presente prautient over a portion of the elannel width cau become negative, that is, back-/low may occur uear the wall which is at rest, and it is seen from Fig. 5.2 that this happens wher P < — 1. In this case the dragging action of the faster layers exerted on fluid par- ticles in the neighbourhood of the wall is insufficient to overcome the influence of the adverse pressure gradiont. This type of Couctte flow with a pressnre gradient, has sorme importance in the hydrodynamic theory of lubrication. The flow in the narrow clearance between journal and bearing is, by and large, identical with Conctte flow with a pressure gradient (c/. Sec. VIc). 2. The Hagen-Py lc theory of flow throngh a pipe. The flow through a straight tube of circular cross-section is the case with rotational symmetry whi corresponds to the preceding case of Wwo-dimensional flow through a chaumel, Let the z-axis be selected along the axis of the pipe, Fig. 1.2, and lot y denote the radial coordinate measnred from the axis outwards. The velocity components in the tangential and radial directions are zcro; the velocity component parallel to the axis, denoted by 1, depends on y alone, and the pressure is constant in every cross- section. Of the three Navier-Stokes equations in cylindrical coordinates, eqns. (3.36), only the one for the axial direction remains, and it simplilies to du) Ldu) _ dp «(Gs +5 y de) dz (5.6) the boundary condition being u = 0 for y = R. The solution of eqn. (5.6) gives the velocity distribution my) = ge SP 9) (6.7) 86 V. Exact solntions of the Navier-Stokes equations: where --dpjdx =: (p, — p,)/- = const is the pressure gradient, to be regarded aa given. Solution (5.7), which was obtained here as an exact, solution of the Navier- Stokos equations, agreca with the solution in eqn. (1.10) which was obtained in an elementary way. The velocity over the cross-section is distributed in the form of a paraboloid of revolution. ‘The maximum velocity on the axis is na ® (- wr) . ” a \~ de The mean velocity %@ = } a, that is, Rt wee (- ar) . (5.8) and the volinne rate of flow becomes Qaamaa3h(_m), (5.9) The laminar flow deseribed by the above solution occurs in practice only as long as the Reynolds number R = @i/v (d= pipe diameter) has a value which is less than the so-called critical Reynolds number, in spite of the fact that the above formulae conatitate an exact solution of the Navicr-Stokes eqations for arbitrary valuos of dpjda, R, and jz, or henee, of i, R, and se, According to experinents y approximately. For R > Rey the flow pattern is entirely different and becomes tur- bulent, We shall disenss this type of flow in greater detail in Chap. XX. The relation between the pressure gradient and the mean velocity of flow is normally represented in enginecring applications by introducing a resistance coefficient of pipe flow, A. Vhis coefficient is defined by sctting the pressure gradient proportional to the dynainic head, i. ¢., to the square of the mean velocity of flow, according to the equation t dp Aes . —fe foe. (5.10) Introdnoing the expression for dp/dx froin eqn. (5.9) we obtain 2d Bpu Bp A= oat i ~ galt that, is o 5 ae (5,11) with (5.12) + This quadralic law which assumes dpjdz ~ @ fits turbulent flow vory woll, It is retained for laminar flow, although in that range dp/dz ~ @, ‘Thus for laminar flow A ceases to bo & constant. a, Parallel flow 87 CTT * 0.255 om dia, a Jae» 0.402 cmaia 4.20 © 0591 cm dia. 010 006 Fig. 6.3. Laminar flow (rough pipes resistance coefficient, A, plotted against 904 Reynolds number (measured by Hagen), from Prandtl-Tietjens 0.02 MELLEL I 700 200 400 600 1000 2000 p-uid ” Hlero R denotes the Roynolds number ealenlated for the pipe diaineter and ancan velocity of flow. The laminar equation for pressure loss in pipes, eqn. (6.11), is in exeellent agreement with experimental resvlis for the laminar range, as secn from Fig. 5.3 which reproduces experimental points moasnred by G. Mugen [10]. From this it is possible to infer that the Magen-Poiseuille parabolic velocity distribution represents a solution of the Navier-Stokes equations which is in agree- ment with experimental results [22]. It is also possible to indicate an cxact solution of the Navier-Stokes equations for the ease of a pipe witl a eireular annular eross- seetion [20]. The problem of laminar and turbulent flow through pipes with exeentric annular cross-sections was discussed theoretically in ref. [38] whieh also contains experimental results. 3. The flow between two concentric rotating cylinders. A further example which loads to @ simple exact solution of the Navier-Stokes equations is afforded by the flow between two concentric rotating eylinders, both of which move at different but steady rotational speeds, We shall denote the inner and outer radii by 7, and rp respectively, and similarly, the two angular velocities by @,, and wa, ‘The Navier- Stokes equations (3.36) for plane polar coordinates reduce to (513) and (5.14) with u denoting the cirenmferontial velocity, The boundary conditions are u =r, for r =r, and w = rg We for r =ry The solution of (5,14) whieh satisfies these re- quirements is 1 tt ryt! [- (0,78 — 0, 4) — 9" (@, —@) | (5.15) u(r) = 2 Fiqnation (5.13) detormines tho radial pressnro distribution reanlting from the motion, 88 V. Exact solutions of the Navior.Stokes equations ‘The case when the inner cylinder is at rest, while the onter eylinder rotates, has some practical significance, In this instance the torque trusmitied by the outer cylinder to the fluid becames 13, (5.16) where h is the height. af the cylinder, ‘Che moment. Mf, with which the fluid acts on the inner eylinder has the same magnitude, ‘Tho arrangement. under cansideration has heen used occasionally for the determination of viscosity. The angullar velocity of the external cylinder and the moment acting an the inner cylinder are measured, so that the viscosity can be evaluated with the aid of eqn. (5.16). We now propose to indicate the velocity distributions in the annnhis between the two cylinders for two particular cases. In Case I, the inner cylinder rotates with the outer ane at rest; in Case I, the inner cylinder dacs not move, hut. the onter enc rotates. Bath flows are called Conette flaw. Denating the ratio of the twa radii hy x = r/re, the width af the anmilus hy s = rz—ry, and the eurrent relative rading by r/re, Fig. 6.4, we find (inner rotating: onter at rest). (5.16) and (inner at rest; anter rotating). (5.161) B Here, 1 == rim is the peripheral velocity of the inner eylinder, and 12 = rave is that for the onter cylinder. Figure 5.4 represents the two velocity distributions in terms af the dimensionless distance from the inner cylinder It is noteworthy that the velocity varies strongly with the ratio x = ri/re of the two radii in Case 1, whereas far Case 11 it is almost. independent of it. When x = n/rz > 1, hoth cages tend to the linear velocity distribution of Coitette flow, as it, occurred between two flat plates in the case represented in Fig. 1.1. The equation of Case IL yields the same limit for r= 0, i, ¢. for x = 0 when no inner cylinder is present. In this case, the fluid rotates inside the auter eylinder as a hody. Hence it. is seen that Case I yields a linear velocity distribution for the two asymptotic eases x = 0 and x = 1. This behavior makes it. casy to understand why the velocity distributions for the other, intermediate values of % differ eo lille from a straight line. In the particular case of a single cylinder rotating in an infinite fluid (r_ > 00, ory O) eqn, (5.15) gives w = 7,2 a/r, and the torque transmitted by the finid to the cylinder becomes M, = 42 yeh r,2 ay. The velocity distribution in the fluid is the meas that around a line vartex af strength Py - 2 ore, int frictionless flow, ar a, Parall 89 - istribution in the aphntus between two, concentric, rotating eyTiaders as ¢ Hh the aid of equs. (5.158, b), a) Case T: inner cytinder rotating; outer cylinder at rest, w2 = 0 b) Case 11: inder at test, on = 0; anter cylinder rotating fy radius oF tuner extinder, ry ~ radline of onter eytinter Tr ar ws : Jt is scon, therefore, that. the ease of frictiouless flow in the neighbourhood of a vortex line constitartes a solution of the Navier-Stokes equations (cf. Sec. 1Vb). In this connexion it may be instructive to mention an example of an exact non- steady sohition of the Navier-Stokes equations, namely that which describes the process of «lecay of a vortex through the action of viscosity. ‘The distrilnttion of the tangential velocity component w with respect to the radial distance r and time ¢ is givert by ulrt) = 3°, {1 — exp (—r%/4 v9} Fig. 5.5. Velocity distribution at varying times in the neighbourhood of a vortex filament caused hy the action of viscosity ‘y= clroutation of tte vortex flament at time (= 0 when viscostty begins to act; yg Tal2 1 Fy 90 V. Exact solutions of the Navior-Stokes equations as derived by ©. W. Oscen [21] and G. Hammel [11]. This velocity distribution is represented graphically in Fig. 6.5 Here J’, denotes the circulation of the vortex filament at time ¢ = 0, i.c. at the moment when viscosity is assumed to begin its action. An experimontal investigation of this process was nndertaken by A. Timme [40}. K. Kirde [17] made an analytic study of the case when the velocity distribution in the vortex differs froin that iinposed by potential theory. 4, The suddenly accelerated plane wall; Stokes’s first problem. We now proceed to calculate some non-steady parallel flows. Since the convective acocleration terms vanish identically, the friction forces interact. with (he local acceleration. The simplest flows of this class occur when motion is started impulsively from rest. We shall hegin with the case of the flow near a flat plate which is siddenly accelerated from rest, aul moves in its own plane with a constant velocity U,. This is one of the pro- bleins whieh were solved hy @. Stokes in his eclebrated memoir on pendulume [35}f- Solecting the x-axis along the wall in the direction of Up, we obtain the simplified Navier-Stokes equation i= byt (5.17) ‘The pressure in the whale space is constant, and the boundary conditions are: <0: u=0~ forally; (5.18) t>0: u=Uy fory=0; u=0 fory oo. The differential equation (5,17) is identical with the equation of heat conduction which deseribes the propagation of hoat: in the space y > 0, when at time t = 0 the wall y == 0 is suddenly heated to a temperature which execeds that in the surround- i ‘The partial differential equation (5.17) can he reduced to an ordinary differ- Jequation by the substitution =e. 5. 1 Vii (5.19) If wo, further, assume us Usiin, (5.20) we obtain the following ordinary differential equation for f (7): P42nfi= (6.21) with the boundary conditions f =| at 1 = 0 and { = 0 at 7 = 00. The solution is us: Uy crfe yn, (5.22) where . oo 2 2 2 rte ny fow (—aP) dy =~ erty = 1 os fox t- py dni . o b + Some authors refer to this problem as the ‘Rayleigh problom’; there is no justification for this designation because the problem can be found fully discnssed and solved in ref. 135). a. Parallel flow o the complementary error Junction, erfe n, has been tabulatedt. The velocity distribu- tion is represented in Fig. 6.6, and it may be noted that the velocity profiles for varying times are ‘similar’, i. ¢., they can be reduced to the same curve by changing the scaic along the axis of ordinates. The complementary error function which appears in eqn. (5.22) has a value of about 0-01 at 7 = 2-0. Taking into account the definition of the thiekness of the boundary layer, 5, we obtain b=2mVriwsyrl. (6.23) It ig seen to be proportional to the square root of the prochict of kinematic viscosity and time. This problem was generalized by FE. Beeker [3] to include more gencral rates of acceleration as well as the cases involving suction or blowing or the effect. of compressihility. Fig. 5.6. Velocity distribution above a suddenly accelerated wall 5. Flow formation in Couette motion. ‘The substilution (6.19) which leads to eqn. (5.21) does not, in general, lend to a solution of the ao-enlied heat conduction equation (5.17) complicatod boundary conditions aro imposed, Sines eqn. (5.17) be obtained by the use of the Laplace transformation and by more direct methods developed in connexion with tho study of the conduction of heat in solids, Many results obtained, c. g., for the temperature variation in an infinite of semi-infinite solid, can be directly transposed and used for the solution of problems in viscous flow. Thus the preceding problem in which the formation of the boundary layer near a suddenly accelerated wall has been investigated can lao bo folved for the case wlien the wall moves in a direction parallel to another flat wall at rest and at a distance h from it. This is the problem of flow formation in Couette motion, i. ¢., Math. Tables + See e.g. Sheppard, “Tho Probability Integral", British Anson. Adv. Sc , New vol. vii (1939) and Works Project Administration “Tables of the Probability Funeti York, 1941, 92 V. Exact solutions of the Navier-Stokes equations the problem of how the velocity profile varies with time Lending asymptotically to the linear distribution shown in Fig. 1.1. Tho differential equation is the same ax before, eqn. (5.17), but with wodified bomdary conditions which now are: $0; u=0 forall y. LOSy shi b> 0: u = Ug for y —0; «=O for yh. .17) which satisfies the boundary and initial conditions can be ‘The solution of eqn. ( ‘en of coutplemontary error functions obtained in the form of a se u Fh = Seren tay — ¥ orfel2 Ot tn = a) wo (5.24) = erfe y — erfe (274 — n) + orf (29, bm) — orfe (Any ~~ m) A orf (4, En) where 1, = h/2 ¥ +i denotes the dinensionloss distance between the two walls. ‘Tho solution is represented in Mig..5.7. ‘The early profiles are atill approximately sitnilar and remam 0, a8 long as the boundary layer has not spread to the stationary wall. ‘The atecceding velocity profiles are no longer “similar” and tend asymptotically to the linear distribution of the stendy state. Flow formation in Conette motion Exact solutions for non-steady Couette flow were derived by J. Steinheucr [33] for the ease when one of the walls is at rest ina stemly flow and is thea suddealy accelerated to a given, constant velocity, ‘To do thia, iL is aecessary wv solve qu. (5.17), which is identioal with the one-dimensional heat conduction equation, by ineans of a Foarier series. A special case in this class of solutions is that when the inoving wall is suddenly stopped so that it represents the decay of Couette flow, 6. Flow in a pipe, starting from rest. ‘The acceleration of a fluid ina pipo is closely related lo the preceiting examples. Suppose that the Muid in an, infinitely long pipe of cireular eross- section is at, rest for ¢ <0. At the instant ¢ = 0 a presstire gradient dpjdr, which is constant In fimo, begins to act along it. The fluid will begin to move tinder the influence of viscous and inertia forces, and the velocity profile will approach asymptotically the parabolic distribution ih Hngen-Poisenille flow. The solution of this problem which leads to a differential equation Invelving Bessel fanetions was given by zymaneki (37[, ‘The velocity profile is druwn in Fig. 5.8 for varions intstants, It is noteworthy that, in the early stages the velority. near the axis is approxiitately constant over the rading and tat viscosily makes itself felt int narrow a, Parallel flow 93 layer near the wall, The influonce of viscosity reaches the pipe contre only in the later stages of motion, and the volovity profilo Lends asymptotically to the parabolic distribution for steady flew. The corresponding solution for an annular cirenlar cross-seotion was givert by W. Mueller [20]. The analogons case when Ue pressure gradiont. is removed instantly was solved hy W. Gorbers [9 ‘The acceleration of u fluid over the whole length of pipe dischssed hero mnat be enrefntly distinguished from the acceleration of a fluid in tho inlet portions of a pipe in steady flow. ‘The rectangular velocity profile which exists in the entrance section js gradually transformed as the fluid progrosses through the pipe with x inercasing, and tends, mder the influence of viscosity, to assume the Hagen-Poisenille parabolic distribution, Since here dujéx + 0 tho flow is not one-dimensional, and the velocity depends om x, ax well as on the radius, This problein wa’ discussed by II. Schlichting [30], who gave the solution for tavo-dimensional flow through aatraight channel, and by L. Schiller [29], and B, Pannis [24] for nsially symmetrical flow (circular pipe): see also Sees, IX} and XID. as Fig. 5.8. Velosity profile in a cirenlar pipe during aceeleration, as given by F, Szymanski (37); 7 = » 1/R® 7. The flow near an oscillating flat plate; Stokes’s second problem. In this section we propose to discuss the flow about an infinite flat wall which cxceutes linear har- monic oscillations parallel to itself and which was first treated hy G. Stokes [35] and later by Lord Rayleigh [25]. Let. x denote the coordinate parallel to the direction of motion and y the coordinate perpendicular to the wall. Owing to the lition of no slip at the wall, the fluid velocity at it must. be equal to that. of the a hy 10 (0,t) = Uy cos nt (5.25) cor wall, Supposing that this notion ix gi (6.17), logebher with the known from the theory we find that, the fluid velocity 2(y,4) is the solution of eq boundary condition (5.25), which, as already mentioned, | conduction, For the ease under consideration u(y,t) == Uy e7'¥ cos (nt—ky) . (5.26) Il is casy to verify that eqn. (5.26) is the required solution if 04 V. Exant solutions of the Navier-Stokes equations Putting n = ky = y Vnj2¥ we have u(y,t) = Ugo" cos (nt — 7). (5.26a) The velocity profile u(y,t) thus has the form of a damped harmonic oscillation, the amplitude of which is Uy o-¥¥*", in which a fluid layer at a distance y has a phase lag y Yn/2¥ with respect to the motion of the wall. Fig. 5.9 represents this motion for several instants of time. Two fluid layers, a distance 2 x/k = 222 y/n apart, oscillate in phase. ‘This listance can be regarded as a kind of wave length of the motion: it is sometimes called the depth of penetration of the viscous wave. ‘The layer which is carried hy the wall has a thickness of the order 8~ V¥/n and decreases for decreasing kinematic viscosity and inercasing frequencyt. Fig. 5.9. Velocity distribution in the neighbourhood of an oscillat: ing wall (Stokes’s second problem) 8. A general class of non-steady solutions. A goneral class of non-steady solutions of the Navier-Stokes equations which possess boundary-layer character is obtained in the special case 1 the velocity components are independent of the longitudinal coordinate, x. The system of equations (3.32), written for plane flow, assimes the forin au, me LN op yy ue - aL dy 7 0 be ay? (5.274) 1 aj ° a (5.27) av ' t ay 78. (5.270) » in eqn. (5,26a) represents also the temperature distribution in the earth which is caused by the periodic fluctuation of tho temperature on the surface, say, from day to day or over Lhe soasons in a yea Ir. b. Other exact solutions 95 If we now prescribe a constant volocity vy <0 at the wall (suction), we notice that eqn. (5.270) in satisfied immediately by a flow for which » =v. and that the pressure p becoines independent of y sirmmultancously. Accordingly, we put — (1/e) (p/éx) == dU jal, where 1 (t) denotes the f streain velocity ata very large distance from tho wall, and hence obtain the following differential ceguation for r(y, #): uy 4, me aU, atu 5 a ay 7 di hae - 28) According to J.P, Stuart [32] thore exints an oxact solution of eqn, (5.28) for the arbitrary external velocity U() ~~ Us EMO). (5.20) ‘This solution ix wy) = Tole + 9% 9) (5.30) where ° (5.31) , 5B Ly) -1—o Substituting the Inst three equations into eqn. (5,28), we are led to a partial differential equation for the unknown function g(y, ¢) = g(7, ¢); this has the form a 0 ag ap ~4 an af(T) +4 ape? (5.32) and the boundary conditions are: n=0:9=0; n= org ‘The following non-dimensional varinbles havo been introduced in the preceding: n= wp (5.33) Solutions of (5.32) have been obtained by J. Watson [41] who employed La formations and who restricted himself to several special forins of the function /(Q). Generally apeaking, the following external flows, U(), have been ineluded: a) damped and undamped oscillations, b) step-like change from one valno of velocity to another, ¢) linear increase from one value to another. In the special case when the external flow ia independent. of time, /(t) — 0, equation (5.32) leads to the simple solution 9(7, 7’) = 0. 'Thia causes the volocily profilo from eqn. (5.30) to hecome identical with the aaympLotic suction profile given later in eqn, (14.6). b. Other exact solutions ‘The preceding examples on one-dimensional flows were very simple, becanse the convective acceleration which renders the equations non-linear vanished identically everywhere. Wo shall now proceed to examine some exact solutions in which these terms are retained, so that non-linear equations will have to he considered. We shall, however, restrict ourselves to steady flows. 9. Stagnation in plane flow (Hiemenz flow). ‘The first simple cxample of this type of flow, represented in Fig, 5.10, is that loading to a slagnation point. in plane, 96 V. Kixact solutions of the Navier-Stokes equations TMT UT f-5.10, Stagnation in plane (low i.e, two-dimensional flow, The velocity distribution in friclionless potential flow in the neighbourhood of the stagnation point at. a =: y =0 is given by Usan; V — ay, where a denotes a constant. ‘This is an example of a plane potential flow which ar- rivos from the y-axis and impinges on a flat wall placed at y = 0, divides into two streains on the wall and leaves in both directions. The viscous flow mnst adhere to tho wall, whereas the potential flow slides along it. In potential flow the pressure is given by Bernoulli's equation. If py denotes the stagnation pressure, and p is the pressure at an arbitrary point, we have in potential flow Po — P= 4 (UU? | V?) = 3 9 a(x? 4-4). For viscons flow, we now make the assumptions weaf(ys 9 =— fy), (5.34) and Po P= Qa? (x? | Fly). (5.35) is way the equation of continuity (4.4¢) is satishie Navier-Stokes equations of plane flow (4.4a,b) are sult tions {(y) and F(y), Substituting eqns. (5.34) and (5. tain two ordinary differential equations for f and F: 1 identically, and the two ent to determine the fune- 5) into eqn, (4.4a,b) we ob- Prat fsa vp (5.36) and [f= hate — » (5.37) b. Other exact solutions 7 ‘The boundary conditions for f and F are obtained from a =: v == 0 at the wall, where 0, and 7 = py at the stagnation point, as well as from « -= U = aa ata large tance from the wall, ‘Thus yenO: f=02 fix 0; FO; promo: fom, Kqus, (5.36) and (5.37) are the two dilferential equations tor h determine the velocity and prossure distributi 6 functions f(y) and Since F(z) docs not. {(y) and then to proceed to find F(y) from the sceond equation. ‘The non-linear differential equation (5.36) cannot be solved in closed terms. In order to solve it numerically it is con- venient to remove the constants a and » by put yorays fy) = Ady). at ABH? — 6g") This ad | Aap”, where the prime now denotes differentiation equation become all identically equal to uni expect to 7. The coelficients of the y if we put a2 AP a2; yA --a2 Wi n= Vou: Hy) = Var $(n). or A Vra; « so that ‘The differential equation for (17) now has the simple form ebb? gL =0 with the boundary conditions yO: b= 0, P=; yoroo: Pool. ‘The velocity component, parallel to the wall becomes ra br #on. ‘The solution of the differential equation (5,39) was first given ina thesis hy K, Hic- menz. [12] and later improved by 1. Howarth [14]. Tt is shown in Fig. 5.11 (see also Table 5,1). The curve $’(7) begins to increase linearly at 7 = 0 and tends asymptotically to unity, At. approximately 1 == 24 wo have ¢' = 0-99, i.e, the final value is reached there with an accuracy of | per cent. If we considor tho corre- sponding distance from the wall, denoted by y = 0, as the boundary Inyer, we have (5.40) 98 V. lxact solutions of the Navier-Slokes equations Tablo 5.1. Functions occurring the solution of plane and axially symmetrical flow with stagnation point, Plane case from L. Howarth [14]; axially symmetrical case from N, Froessling [8] plano axially symmetrical a dg ou ay ie Za ay n= Voy, 8 lata] ape [YRS 2] ¢ “age 0 0 0 12326 0 0 13120 02 0-0233 02266 1.0345, 2 0127 11705, O4 0-0881 04145 08463 4 00487 10298 06 01867 05663 06752, 0-6 01064 0-8910 O8 0-31.24 0-6859 0-251 O8 0-1799, 0-7563, 10 00-4592 0-7779 0:3980 1-0 02695 06283 1-2 06220 00-8467 02938 12 O3TLT 05097 14 0-7967 08968 O20 14 04841 0-4031 1-6 09798 | 0-9323 | 0-1474 16 06046 03100 18 11689 | 09568 | 01000 18 07313. 2315 20 13620 0-9732 0-0658 20 08627 0-1676 22 1-5578 0-9839 0-0420 22 0-9974 OALTS: 24 1-753 0-9905, 0-0260 24 41-1346 00798 26 41-9638 09946, 0-0156 26 12733 090523, 2-8 24530 0-9970, 00090, 28 14131 00-0331 30 2-3526 0-984 00-0051 30 1-5536 0-0202 32 2-5523 0-9992 | 00028 32 16944 0-0120 34 2-7522 0-9906, 00014 34 18356, 09-0068. 36 2-9521 00-9998 | 0-0007 36 1-9769 00037, 3-8 31521 09999 00004 38 22-1182 00020 40 33521 1-0000 00-0002 40 2-2596 0-0010 42 36521 41-0000 0-0001 42 24010 0-0006 44 37521 1.0000 0-0000 44 2.5423 0-0003, 46 3-952 1-0000, 09-0000 46 2-6837, 0-0001 Hence again, as before, the layer which is influenced by viscosity is small at low kinematic, viscositics and proportional to ¥», The pressure gradiont ap/@y becomes proportional to g@Yva and is also very small for small kinematic viscosities. It is, further, worth noting that the dimensionless velocity distribution u/U and the boundary-layer thickness from eqn. (5.40) are independent of x, i. ¢., they do not vary along the wall. ' Tho type of flow under consideration does not occur near a plane wall only, but also in two-limensional flow past. any cylindrical body, provided that it has a blunt nose near the stagnation point, In such eases the sohition is valid for a amall neigh- bourhood of the stagnation point, if the portion of the curved surface can be replaced by its tangent plane near the stagnation point. b, Other exact solutions 99 The non-steady flow pattern whieh results upon the superposition of an arbi- trary, Lime-dependent transverse motion of the plane was sludied by J. Watson [42]. ‘The special ease of a harmonic transverse motion was solved carlior by M. B. Glauert. (14] in Chap. XV). 9a. Two-dimensional non-steady stagnation flow. Tho carc of non-stondy, two-limonsionat flow studied by N. Rott [28a] conatitutos generalization of Uhe preceding case, We consider the ease of two-dimensional stagnation flow depicted in Fig. 5.10 and bounded by » wall at y We asaimo that the velonity at x large distaneo from the wall is dlireetod towarda the wall, snd that the wall itself performs a harmonic motion in its own plano. In the resulting flow patter the velooity remains steady ata largo distance (y —-co), whereas near the wall it acquires a non. steady pattern of the same kind as that near the oscillating wall of Fig. 5.0 (Stokes's second problent). Aocording to [28a], it is possible to integrate the nonstendy Navier-Stokes equation (44a, b, e) by asutming u(z, yt) =axd(y) 1 6 g(y) exp (int) (5.40a) oy) = -- (a v8 O(n), the aaine way as waa done in eqn. (5.34). As far as the proaure ia concorned, we put P= po — (‘/2) 9 a? 2? - gvaF(y). (6.400) Here, 1 = y(a/v)!!? denotes the ditnensionless dlistance from the wall from eqtt. (6.38), is the constant ainplitude of the wall oscillating in ite own plano, and « is the cirenlar frequency of this oscillation, ‘The preceding assumptions (6.402, h, ¢) are introduced into the Navier-Stokes equations (44a, b, ¢), and the problem is reduced to solving the following system of equations: er ger — bad . (5.40d) go +9 b~g(b Vik) =0, (5.406) Bp mF bY. (5.401) Here k = w/a denotes the dimonsionleas frequency of tho wall oscillation. The differential equa- tions (640d) and (5.40e) result from the non-steady Navier-Stokes equation in the x-direction, eqn. (4.4), when Uso velocity component 1 is reprosentod as the am of a alendy tern, ¢.', and an unsteady term, g, as was done in eqn. (5.40a), ‘Tho function #(1) satistios the boundary eondi- tions # (0) = 6°(0) D and ¥°( A comparison betwoen eqns. (6.39) and (6.40d) shows that this funetion ix well-known solution of the steady-rlato problem «duc to Hiemenz. boundary conditions cnticnl with the = 0. 90) =1 and gf It is seon from eqns. (5.404) and (5.40¢) that in this oaso the atcady component is independent of the superimposed non-steady component. The differential equation (6.400) for the nou-stendy contribution g of the x-component of the volocity can be easily solved, becanee the funetion 4(n), Table 6.1, is known. Further details concerning this problem oan be found in [28n}. ‘The reader may also consult the papers by M. Glauort, [14] in Chap. XV, and J. Watson, [65} in Chap. Xv. Foo V. Exact solutions of the Navier-Stokes equations: 10. Stagnation in three-dimensional flow. In a similar way it is possible to obtain an exact, solution of the Navier-Stokes equations for the threc-dimensional ease. of flow with stagnation, i.c,, for the axisymmetrical case. A fluid stream impinges on a wall at right angles to it and flows away radially in all directions. Such a case oceurs in the neighbourhood of a stagnation poit of a body of revolution in a flow parallel to its axis. 16 4 12 10 08 06 a4 Fig. 5.1). Velocity distribution of 02 plano and rotationally flow at a stagnation point of yinnietrical ‘Yo solve the problem we shall use cylindrical coordinates r, 4, z, and we shall assume that the wall is at z = 0, the stagnation point is at the origin and that the flow is in the direction of the negative z-axis. We shall denote the radial and axial components in frictionless flow by U and W respectively, whereas those in viscous flow will be denoted by u == u(r,z), and w = w(r,z). In accordance with eqn. (3.36) the Navier-Stokes equation for rotational symmetry can be written as au ou 1 ap ‘au, 1 du ou, ow wa ba = 6 & +o (Se re +3), aw ao 1 ap, | (ate, 1 aw , ate 5.41 a i oat (ur or a), 641) because rg ~~ 0 and aj -~ 0, anid we have pub r, =e and == t0. hb. Other exact solutions 101 The boundary conditions are z=0: u=0, w-0; zocor wil, For the frictionless case we can write Uoour; Wo .2az, (5.42) where a is a constant. Tt is scen at once Ural euch a solution satisties Lhe equation of continuity. Denoting once more the stagnation pressure by pig, we find the pressure in ideal flow: Po — P= 4 (U2 EW) = 4 gar? | 42%). In the case of viscous flow we assume the following form of the solutions far the velocity and prossure disteibutions war f(z): wee 2 f(z). (5.43) Po — P= Fa? |r| FGM. (44) It can be casily verified that a sohition of the form (5.43) satisfies the equation of continuity identically, whereas the equations of motion lead to the following two equations for f(z) and F(z): 25 =a eof", (5.45) 2s = bah’ — vf". (6.46) The boundary conditions for f(z) and F(z) follow from eqn. (5.412), and are I a As before, the first of the two equations for f and F can he freed of the constants ¢? and y by a similarity transformation, which is identical with that. in the plane case, thus I (2) = Var $2). The differential equation for $(2) simplifies to Po BGP — PBA Ld (5.47) with Ute bomdary conditions f=0: be P=, C ‘The solution of eqn. (5.47) was first given by F. Tomann [13] in Ute farm of a power series, The plot of ¢' = u/U/ is given in Fig. 5.11 together w the values for ¢ given in ‘Table 5.1 have been taken from a paper by N. Froessling {8}. loz V. Exact solutions of the Navier-Stokes equations: IL. Flow near a rotating disk. A further example of au exact solution of tho Navier-Stokes equations ix furnished hy the flow around a flat disk which rotate about an axis perpendicular to its plane with a uniform angular velocity, w, ina fluid otherwise at. rest. ‘The layor near the disk is cartied by it through friction and is thrown outwards owing to the action of centrifugal forces. ‘This is compensated by particles which flow in an axial direction towards the disk to be in turn carried and ejected contrifugally. ‘Thus the case is seen to be one of fully throe-dimensional flow, i.c¢., there exist velocity components in the radial direction, 7, the cireum- ferential direction, g, and the axial direction, 2, which we shall denote respectively by tt, ». and w, Aut axonometrie representation of this flow field is shown in Fig. 6-12. AL first the calculation will be performed for the case of an infinite rotating plane. It will then be casy to extend the-result to include a disk of finite diameter D == 2 R, on condition that, the edge effect is neglected. ‘Taking into account rotational symmetry as well as the notation for the problem we can write down the Navier-Stokes equations (3.36) as: ae tye Or Be (5.48) Fig. 5.12. Flow in tho noighbour- hood of a disk rotating in a fluid at rest, Velocity components: u-radial, »-cireum- fercritial, w-axtal, A iayer of fuid Is car- ried by the disk owing, to the action of vineous forces. The centrifugal forees in the Uhta iayee give rise to soconttary flow which 1 directed radially outward b. Other exact solntions: 103 ‘The no-slip condition at the wall gives the following boundary conditions: z=0: u=0, | za0o: u=0, y 64) We shall bogin by estimating the thicknoss, 5, of the layer of fluid ‘carried’ by the disk [23]. Lt is olear that the thickness of the layor of fluid which rotates with the disk owing to friction deereases with the viseosity and this view is confirmed when compared with the results of the preceding examples. ‘The centrifugal force per unit volume which acts on a fluid particle in the rotating layor aba distance r from the axis is equal lo gr w2. Henee for a volume of area dr -ds and height, 3, the centri- fugal forec becomes: g rw? 6dr ds. The saine nt of fhiid is acted upon by i shearing stress t,,, pointing in the direction in which the fluid is slipping, and forming an angle, say 0, with the cireumferential velocity. ‘The radial component of the shearing stress must now be equal to the centrifugal force, and hence Ty sin Oris = 970? 5 drde or Ty sin 0 = e@rwd. On the other hand the circumferential component of the shearing stross must be proportional to the velocity gradient of the cirounferential velocity at the wall. This condition gives 1 C8 OW per w/d Eliminating 1, from these two equations we obtain ~~ ” tand. o If it is assumed that the direction of slip in the flow near the wall is independent of the radius, the thickness of the layor carricd| by the disk becomes r aw /Z, which is identical with the result obtained in tle case of the oscillating wall on p. 94. Further, we can write for the shearing stress at the wall tw~werad~erayra. The torque, whiel is equal to the product of shearing stress al Une wall, area and arm beeomes M~r1,R~e Rw ro, (5.50) R denoting the radius of the disk. In order to intograte the system of equations (5.48) it is convenient to introduce a dimensionless distance from the wall, ¢-~ 2/3, thus putting (5.51) 104 V. Exact solutions of the Navier-Stokes eqnations. Farther, the following assumptions are maile for the velovity componeuts ant pres- sure u=roF(t); v=raQt); w= Vr@ Me) (5.52) P= P(2)=erw PC). Inserting these equations into eqns. (5.48) we obtain a system of four smultancous ordinary differential equations for the fanetions F,G, 11, and P: 2F4-H' P4PHO @aF’ 2FG+HE—@’ P+ HH —H" = ‘The boundary conditions ean be calemlated from equ. (5.49) and are: t=0: F=0, G@=1, N=0, t=co: F=0, G=0. ‘The first sohition of the system of eqns. (5.53) by an approximate metho was given by a method of nnmerjeal integration}. They are plotted in Fig. 5.13. ‘The sta valnes of the solution indicated in Table 5.2 were given by ©. M, Sparrow and J.L. Gregg [32]. 0 08 a6 04 02 Fig. 5.13, Vek distribution nearadisk rotating in a Mid at rest, point, the velocity field i is “the. first to be. evaluated from the equation of ‘cont and the equations of motion parallel to the wall. The pressure distribution is found subsequently from the equation of motion perpendicular to the wall. + This solution was obtained in the form of a power series near ¢ = 0 and an asymptotic series for large values of £ which were then joined together for moderate values of b. Other exact solutions 105, ble 6.2. Values of the fu ions necded for the description of the flow of a disk rotating in a at rest. calewlated at the wall and ata large distance from the wall, as ealeulated by B.M. Sparcow and J. L. Grog [32] F | @ | u | P 0 0-510 0-6159 0 0 oo 0 0 0-8845 0-3912 Itiescen from Fig. 5.18 Unat the distance from the wall over whieh the periphersd velocity is reduced to half the disk velocity is do. © Vr/ar. It is to be noted from the solution that when 5 « Yx/o is small, the velocity components w and 2 have appreciable values only ina thin layer of thickness y/o. The velocity component 1, al any rate, small and of the order frar. The inclination normal to the disk i of the relative str if the wall is imagined at rest. and the fluid is taken to rotate at a | from the wall, becomes _ _ (oujaz) FM) _ 51 tan fo = — (Bi), = — 5g = Ba by = 396°. Although the calculation is, strictly speaking, applicable to an infinite disk only we may utilize the same results for a finite disk, provided that its radius RF is lar; compared with the thickness 6 of the layer carried with the disk, We shall now evaluate the turning moment. of such a disk. ‘The contribution of an annular disk element. of width dr on radius r is dM = —22rdrrt,g, and hence the moment for a disk wetted on one side becomes R M = —2n [rr tgdr. é Here t.g = 2(2v/d2)q denotes the circumferential component of the shgarintg stress. From cq. (5.52) we obtain Teg = OP NE en2" (0). Ilenee the moment for a disk wetted on both sides lecomes: 2M = —xQ Ro)? G’ (0) = 0616 x 9 R4(vw3)!”, (5.54) It is customary to introduce the following dimensionless moment coefficient, (5.55) 106 V. Exact solutions of the Ni Stokes equations This gives ___ 2x@(oyv'? Cu = i or, dofining a Reynolds number based on the radins and tip velocity, aud introducing the numerical value — 27 @"(0) = 3-87, we obtain finally C Vi (5.56) Fig. 5.14 shows a plot of this equation, ourve (1), and compares it, with measure- ments [39]. For Reynolds numbers up to about R =3 x 105 there is excellent agreement between theory and oxporiment. At higher Reynolds numbers the flow beeomes turbulent, and the respective case is considered in Chap. XXI. Curves (2) and (3) in Fig. 5.14 aro obtainod from the turbulont flow theory. Older measurements, carried out by G. Kempf [16] and W. Schmidt, [31], show tolerable agreement, witlt Lkcorolical results, Prior to these solitions, D. Riabouchinsky [26], [27] established empirical formulae for the turning moment of rotating disks which wero based on vory careful measurements. These formulae showed very good agreoment with the theorotical equations discovered subscquently. ‘The quantity of liquid whieh is pumpod outwards as a result of the centrifuging action on the one side of a disk of radius R is Q=20R fude. z=0 "e LIT Ty 20 141 ft 100« © NACA Report We. 793 oH so Mach number 40 00240062 = Y 0.48 to 1.69 ao * Kempt rTP - © WSchmic . 20 °WSchmidt |_| Fig. 6.14. Turning mo- ment on a rotating disk; eurve (1) from eqn. (5.56), laminar; curves (2) and (3) from eqns. =~4 (21.30) and (21.33). tur- bulent b. Other exact solutions: 107 Calculation shows that Q = 0-885 x Ry @ = 0-885 x R83 a R-V?. (5.57) ‘The quantity of fluid flowing towards the disk in the axial direction is of equal magnitude. Ft is, further, worthy of note that the pressure difference over the layer carrie by the disk is of the order gy, i. e., very sinall for small viseosilic pressure distribution depends only on the distance from the wall, and the radial pressure gradicnt. A generalised form of the preceding problem has been studied by M. G. Rogers and G.N. Lance [28] who assumed that the fluid moves with an angular velocit Q = sw at infinity, With this assuinption, the second equation (5.53) becomes modificd to PEP ~G?— Fs? =0, and the second boundary condition for the function @(2) must be replaced by G(co) = 8. In this connexion a comparison should be made with the case of rotating flow over a fixed disk given in See, XIa. Numerical solutions for rotation in the saine sense (¢ > 0) ean be found in [26], When the rotations are in opposite senses (s<0), physically meaningful solutions can be obtained for s < — 0-2 only if untform sttelion at right angles to the disk is admitted, The problem of a rotating disk in a housing is discussed in Chap, XXL. Iv is particularly noteworthy that the solution for the rotating disk as well as the solutions obtained for the flow with stagnation are, in the first place, exact solutions of the Navier-Stokes equations and, in the second, that they are of a boundary-layer type, in Lhe sense discussed in the preceding chapter, In the limiting case of very small viscosity these solutions show that. the influence of viscosity extends over a very small layer in Uke neighbourhood of the solid wall, where in the whole of the remaining rogion tho flow is, practically speaking, identical with the corresponding ideal (potential) case. ‘These examples show further that the boundary layer has a thicknoss of the order ¥'¥ + ‘The one-dimensional examples of flow discussed previonsly display the same boundary-layer character. In this connexion the reader may wish to consult a paper by G. K. Batchelor [2] wh discusses the solution of the Navier-Stokes equations for the case of two co-axial, rotating disks placed at, a certain distance apart, as well asa paper by K, Stowartson [34]. An extension of the preceding solution to the case of uniform suction is due to J.T. Stuart {92} in Chap, XIV) and to I. M. Sparrow and dL. Gregg (see p. 3 in [32]). The latter contains also an analysis of the case with honiogencons blowing. ‘The limiting easo of very vigorous blowing was disenssed by IL. K. Kuikest [18 12. Flow in convergent and divergent channels. A further class of exact solutions of the Navier-Stokes equations can he obtained in the following way: Let it be assumed that the fainily of straight lines passing through a point in a plaue constitute the str 08 of a flow. Let the velocity differ from line to line, which incans that it is assinted to bo a function of the polar angle g. The rays along which the velocity vanishes can then be regarded as the solid walls of a convergent or a divergent chatnel, The coutinuity equations can be satisfied by assuining that the velocity along every ray is inversely jtroportional to the ¢ from the or the radial velocity x tins the forth a ~ F(f)/r, or, if F ia to be di 108 V. Exact solutions of the Navier-Stokes equations: ua 2 Fi). ; The peripheral velocity vanishes everywhere. Introducing this form into the Navier-Stokes equations written int polar coordinates, eqn. (3.36), and eliminating, pressure from the equa in the r and ¢ directions, we obtain the following ordinary ditlerentinl equation for Fg): 2PM LAR 4 BO, Integrating once, we are led to the equation PP4aP EP 4K =O. (5.58) ‘The constant K denotes the radial pressure gradiont at the wall, K = —-(I/o) (@p/ar) (r/), where we have I = Oforg = xandg — —a.ns wellas F’ = 0 forg = 0. The solution of eqn. (5.58) was given by G. Hamel [11}.'Tho fariction F ean be expressed explicitly as an elliptic function of. We shall now briefly sketch the character, of the solution refraining from discussing the details of the derivation. ‘The graph in Fig. 5.18 shows a family of velocity profiles for a con- vergent, anda diyorgent chanitel for different. Reynolds numbers plotted on the basis of the nomerical calculations performed by K. Millsapsand K. Pohthauser {19}. The velocity distribution for the convergent, and for the divergent channel differ markecly from each other. In the latter case, they also differ markedly for different Reynolds nanthers. In a convergent channel the velocity distribution for the highest Reynolds nunther (R= 5000) reniains nearly constant over a large centre-portion and decreases steeply to zero near tho walls; thus it exhibits in this case a clear “boundary-layer character. Ina divergent channel the shape of the velocity profiles is markedly affected by the Reynolds numbor. Bach of these volocity distributions is more curved at the centerline than tite parnhola that. characterizes flow through a channel with parallel walls, The velocity distribution for (he largest Reynolds number, curve 7, is distinguished by the fact that it shows two regions of back. flow. ‘Thus, the velocity vanishes at four points, Since the wall could place itsell at any onc of these points, it is possible to envisage Uris velocity distribution at am ineluded angle of 10° with two symmetric rogions of back-flow or at an included angle of 0-9" but with a single, asymmetric region of back-llow, Such asymmetric velocity distributions are actually observed, and the back- flow signals the start of aeparation. Fig. 5.15. Velocity distribution in a convergent and a divergent. channel after G, Mantel [11] and KC, Millsaps and K, Pohthansen [19] angle 2a WP Reynolds number R= m4 7/" Convergent ehannet Dive Curve t: R= 5000 Curve 5: = 684 Curve 2: R= 1342 Curve 6: R = 1942 Carve 3: R= at Curve 72 = 5000 Curve 4 refers to a channel with parallet walls {Poiseullic’s paroholie velocity distribu of Fig. 5.1) b. Other exact solutions 109 In the papor referred to above, (i. Hamel has set himself the problein of caloulating all three-dimonsiortal flows whose streamlines are identical with those of a potential flow. The solution consisted of streamlines in the shape of logarithmic spirals. The case of radial ow considered lere, and the case of potential vortex-flow, discussed in See, V 3, constitute parti- onlar examplos of this genoral solution. ‘The preceding example of an exact rohition exhibits once more the boundary-layer character of the flow. In partionlar, in the caso of a convergent channel, Ute existence of m thin lay Uhe wall is confirmed together with the fact that the influence of viscosity is concentrated in . Further, the calculation eonfirms that the boundary-layer thickness increases as ¥ ¥ here too. ‘The divergent case exhibita an additional phenomenon, that of back-flow, and, resulting from it, separation, This is an esscntial property of all boundary-layer flows and we sball discuss’ it later in greater detail on the basis of Ue equations of boundary-layer flow. Its existence is fully confirmed by experiment. ‘The cases of two-dinensional and axi-symmetrical flow through channels with small angles of divergence have been investigated carlier by H. Blasius [5] froin first, principles, i. c., with the aid of the Navier-Stokes equations. In this connexion it was shown that laminar flow ean support only a very amtall prosaure increase without the incidence of separation. The condition for the avoidance of back-flow at the wall in a divergent tube of radius 2(x) was found to be dRjdz <12/R. (condition for separation), where R= id/y denotes the Reynolds number referred to the mean velocity of flow through the channel and to its diamieter. Th more modertt times M. Abramowitz [1] extended these calculations for divergent channels, ani discovered that. the point of separation itoves downstream from the chatnel entrance as the Reynolds humber is increased and as the angle of divergence is decrensed. 13. Concluding remark. This example coneludes the disenasion of exact solution of tho Navier-Stokes eqnations and the next topic will deal with approximate solutions. In the previous description an exact solution meant a solution of the Navier-Sto- kes equations in which all its terms were taken into account, provided that they did not vanish identically for the problem. In the following chapter we shall concern ourselves with approximate solutions of the Navier-Stokes equations, that is, wilh solutions which are obtained when small terms are neglected in the differential equa- tions theinselves, As alrealy mentioned in Chap. IV, the two limiting cases of very large and very small viscosity are of particular jinportance. In very slow, or so- called creeping motion, viscous forces are very large compared with inertia forces, and in boundary-layer inotion they are very small. Whereas in the first case it permissible lo omit the inertia terms completely, no such simplification is possible in houndary-layer theory, because if the viseous terms arc simply disregarded the physically casential condition of no slip al. the solid boundary camiol. be mot. K. W. Manglor [6] developed a general theory for the solution of tho Navier-Stokes equations for the case of two-dimensional laminar flow at very high Reynolds numbers, that is for lows in which the effect of visvosity is included, and which possess boundary-layer character, In Pranctt’s boundary-layer theory (sce also Chap, VII for details) the contour of the solid body in the stroam is preaoribed and the effect of viscosity ia accounted for only in the thin layer adjoining (he wall, By contrast, the new procedure is an indirect one, Instead of the contour of tie reat body, the thoory prescribes an appropriate forin for the so-called displacement contour which surrounds the body. The displacement contonr makes an allowance fot the displacement effect. exerted on the external flow and on the wake, This permits us to detornine the external, frictionless flow about the displacement contour; the next step consists in the compritation of the flow field in the {rictional layer with the ald of an asymptotic treatment of the Navier-Stokes equations for very large Reynolds numbera which yiolds, finally, the real shape of the body. The remarkable feature of this new procedure consists in the fact that the borndary-layer calentation can be carried beyond the point of separation. This is in contrast with Prandia boundary-layer theory which 110 \. Exact sohitions of tle Navicr-Stokes equations can be applicd at most ax far as the point of anparation only, Furthermore, the new theory sne- ccods in some caaca even with the evaluation of the complex flow patterns which exist in the region of back-flow beliind the point of separation as woll as that. in the region of re-attachment. ve theory to that given by Prandtl must suffice here. ‘The ‘The brief moution of an alternal emaindor of tis hook is baaed on Prandtt’s tine of honvdary-layer Ueory expounded in the Mtonght, References [1] Abramowitz, M.: On backflow of m viscous finid in a diverging chanel. J. Math. Phys, 28, 1-2) (1949 [2} Batchelor, stendy uon-rotationally synumetri« flow. Quart. J. M [3] Becker, E.: Eine einfache Verallgemeincrung der Raytcigh 152 (1960). [4] Borker, R.: Integration des équations du mouvement dun fluide visqueux incompressible. Contribution to: Handbuch der Physik (S. Fliigge, od.) VI1//2, 1-384, Berlin, 1963. [5] Blasius, H.: Laminare Strémunng in Kattilen wechscuder Breite. Z. Math. u, Physik 58, 225 (E910). [6] Uatherall, P., and Mangler, K.W.: The integration of the two-dintensioal laminar bonudary- Jayer equations past the point of vanishing skin friction. JFM 26, 163—182 (1960). [7] Coctrau, W. G.: The flow due to a rotating disk. Proe. Cambr, Phil, Soc. 30, 365-375 (1034). [7a] Florent, P. aud Pende, J.L.: Beoulement taminaire d'un fide visquenx incompressible entre deux disques poreux. J. Méeanique 14, 436-- 459 (1975). [8] Frdssting, N.: Verdunstung, Wirmoiibertragung und Geschwindigkeitavertoilung bei dimensionaler und rotationssyminctriscler laminarer Grenzsehichtatronumg. Lunds, Ui Arsekr. N. F. Afd. 2, 35, No. 4 (1940). (9) Gerbora, W.: Zar inataliondren, lantinaren Strémnng ciner inkontpressiblen zhen Fliissig- keit in kreiszytindrischon Roliren. 7% angow. Physik 3, 267—271 (1951). [0] Hagen, G.: Uber die Rewegnng des Wassers in ongen zylindrischon Rohren. Pogg. Aun. 46, 423— 442 (1339). U1] Hamol, G.: Spiralformige Bewegung ziher Fhissigkeiten, Jabresbe Vereinignng 25, 34-60 (1916). [12] Hicmenz, Ko: Die Grenzschicht an cinom In don gleichformigen tauchton geraden Kreiszylinder. Thesis Gottingen 1911. Dingl. Polytech. J.326,321 (1911), {13} Homann, #.: Der Kintinsa grosser Ziihigkeit, bei der Stromung um den Zylinder und am die Kngel. ZAMM 16, 153—164 (1936); Forseltg. Ing,-Wes. 7, I-10 (1936). [14] Howarth, L.: On the ealentation of the steady flow in the boundary layer near the surface of n cylinder in a stream. ARC RM 1632 (193! [15] von Kérmdén, The: Ober laminare und tirbulente Reibung. ZAMM /, 23 NACA ‘TM 1092 (1946); see also: Coll, Works 17, 70-97. U6] Kempf, G.: Ober Reibungswiderstand rotierender Schoiben. Vortriige auf dem Col Hydro. nnd Acrodynamik, Innsbruck Congr. 1922; Borlin, 1924, 108. [17] Kirde, K.: Untoranclumgen iibor die zeitliche Weiterentwicklung cites Wirbela mit vor- gegohoner Anfangavertoilung. Ing.-Arch, 3/, 385 - 404 (1962), [18a] Mettor, G.L., Chapple, P. J. and Stokes, V.K.: On the flow betweon a rotating and a stationary disk. JIM 3/, 95— LEZ (1968). 118] Kuiken, H.K.: The effect of normal blowing on the flow near a rotating disk of infinite extont. FFM 47, 789--798 (1971). " [19] Mitisapa, K., and Pohthansen, K.: Thermal distriution in Joffrey-Hamet flows between nonparaltel plano walls, JAS 20, 187-- 196 (1953). {20] Miitler, W.: Zum Problem dor Antaufstrdmung citer Missigkcit im geraden Robr mit Kreis- ting. und Kroinquerselititt. ZAMM (6, 227—238 (1 [21] Oscen, (.\W.: Ark. £ Math. Astron. och. ys. 7 (1911); Hydromechanik, Leipzig. 1927. p. 82. [22] Poisenitte, 1.: Rechorehes expérimentolles sur le mouventent des Tiquides dana les tubes de (réa polite diamétres, Comptes Rendus 1, 961 --967 and 1041-1048 (1840); 72, 112 (1841)5 in more dotnil: Memoires dea Savants Blrangers 9 (1846). K.: Note ont a class of aolutions of the Navier-Stokes equations representing Appl. Math. 4, 29-41 (1951). onzschielt. ZAMP J7, 146 — Dt, Mathematiker- insigkcitsstromt singe- ~ 252 (1921); jot der References ML [23] Prandtl, L.s Fithrer durch dic Strénungelohre. 6th ed., 500, 1985; Kngl. transl, Blackie and Son, London, 1952. [24] Punnia, B.: Zur Bereohinung der laminaren Eintoufstromung iin Rohr, Diss. [25] Rayleigh, Lord: On the motion of olid bodies through viscous liquid. Ph TIT (1911)3 also Sch. Papers VE, 29. [26) Riabouehi Bull. de PIustitut Acrodyn. de Kontehino, 5, 5 —34 Moscow (1914)s see also. Roy. 40 348 ancl 377--379 (1935). ; [27) Riabonet : tance de frotlomont. des disques tonrnant dans in tide et {es squations iitegrales appliquées & co probléme. Comptes Rencus 233, 899-901 (1951). . and Shipman, J. 8.: Computing of the flow botwoon a rotating and k, JIM 73, 53-~03 (1976). : aa , ONe Me rofationutly aymumetric flow of a viscous fluid in the rotating disk. JIM 7, 617—631 (1960). ty of a slagintion point. Quart. Appl. Math. ingen 1947. . Mag. 27, 197 — presence of at {28a} Rott, N.: Unsteady viscous tlow itt the vi 13, 51 (1055/50). Untersuohiungon fiber laminare und turbulente Strémung. VDI-Forschungs- heeft, 248 (1922). [29a] Schohoiri, M.P.: Nithorungslésungen dor Navier-Stokes’schon Dilferen aweidimensionale stationire Laminarstrimung konstanter Viskositit in konvexen konkaven Diffusoren und Diisen. ZAMP 27, 9—21 (1976), (30) Scliichting, H.: Laminare Kanaleitlaufstrémung, ZAMM 14, 368—373 (1934). [31] Schanidt, We: faclios Mesaverfattron fiir Drohmomtento, Z. VDI 65, 441 —444 (1921), [32] Sparrow, J.M., and Gregg, J,.L.: Mass transfer, flow and tent transfor about a rotating disk. Transactions ASME, J. Heat Tranafer 82, 294—302 (1960). : [33) Stoinheuer, J.: Kine exakte Lisung der instationiren Conette-Stromung. Proc. Seiontili Soe. of Braunschweig XVII, 164—164 (1968). [34) Stewartson, K.: On the flow between two rotating coaxial disks. Proc. Cambr. Phil, Soc. 49, 333-—341 (1063). [35] Stokes, G. G.: On the effect of the internal frietion of fluids on the motion of pendulums. Cambr, Phil, Trans, 1X, 8 (1853); Math. and Phys, Papors, Cainbridge, 117, I —141 (1901). [36] Stuart, J.T.: A solution of the Navior-Stokes nud ottergy equations illustrating the reapouse of skin friction and temperature of au infinite plate thermonieter to fhictuations in the stream velocity. Proc, Roy. Soc, London A 23/, 116—130 (1955). [37] Szymanski, F.: Quelques solutions exactes des équations de Vhydrodynamique de tluide visquoux dans te ces d'un tube eylindrique. J. de math. pures ct appliquécs, Series 9, 11, 67 (1932); see also Proc. Intern. Congr. Appl. Mech. Stocktotin 7, 249 (£930). [38] Tao, L.N., and Donovan, W.F.: Through-flow in concontric and exoentzic annul of fine clenrance with and without relative motion of the boundaries, Trans. ASME 77, 1291-- 1301 (1065), [39] Theodorsen, Th., and Regier, A.: Experiments on drag of revolving disos, cylinders, and streamline rods at high speeds, NACA Rep, 793 (1944). [40) Timme, A.: Uber die Gesohwindigkeitsverteitung in Wirbeln, Ing.-Arch. 25, 205226 (1957). {41} Watson, J.: A solution of the Navier-Stokes equations illuatrating the response of a laininar boundary layer to a given change in tho exterun! stream velovity, Quart. J. Moch. Appl. Math, 77, 302 ~325 (1958). [42] Watson, J.: Tho two-dimonsional laminar flow noar tho stagnation poiut of a cylinder whieh has au arbitrary tranaverso motion, Quart, J. Mooh, Appl Math, 12, 1765-~1h0 (159). CHAPTER VI Very slow motion a, The differential equations for the case of very slow motion ir this chaptor we propose (o risers some approximate solutions of the Navier- totes oxytations whieh are valid tn the Hiniidag enso when the viscous for considerably greater than the inertia forers. Since the inertia forces are proportional to the square of the velocity whercas the viseous forces are only proportional to its first, power, it is easy to appreciate that a flow for which viscous forces are dominant. is obtained when the velocity is very small, or, speaking more generally, when the Reynolds number is very small, When the inertia terms are simply omitted from the equations of motion the resulting solutions are valid approximately for R < 1. This fact can also be deduced from the dimensionless form of the Navier-Stokes equations, eqns. (4.2), whore the inertia terms are seen to be multiplied by a factor R = @ V Ue compared with the viscous terms. In this conuexion we may remark that in each particular case it is necessary to examine in detail the quautities with which this Reynolds number is to be formed. However, apart from some special cases, motions at. very low Reynolds numbers, sometimes also called creeping motions, do not occur too often in practical applications}. It is seen from eqns. (3.34) that when the inertia terms are neglected the incom- pressible Navier-Stokes equations assume the form grad p =p V2w, (6.1) divw =0, (6.2) or, in extended form (6.3) + In tho case of @ sphere falling in air ( . 160 x 10 © £%/s0c) wo obtain e.g. R when the dinineter d = 0-04 in ( 0-00333 {t) and the velocity V = 0-048 ft/sec. b. Paralicl flow past a aphero 3 This system of equations must be supplemented with the same boundary conditions as the full Navier-Stokes equations, namely those expressing the absence of slip in the fluid at the walls, ie. the vanishing of the normal and tangential components of velocity: My = 0, % 220 at walls. (6.5) An important characteristic of creeping motion can be obtained at ouce from oqn. (6.1), when the divergence of both sides is formed and whon it is noticed that the operations div and V7? on the right-hand side may be performed in the reverse order, ‘Thus, with eqn, (0.2) we have div grad p = V2 p -0. (0.0) ‘The presaure field in creeping motian satisfies the potential equation and the jiressure p(«,y,2) is a potential function, ‘The equations for two-dimensional creeping motion become particularly simple in form with the introduction of the stream function p defined hy «= dy/dy and v = — Ayam, As explained in Chap. IV, and as scen from equs. (6.3), when pressnre is cfiminated from the first two equations, the stream fanclion must satisfy the equation Vip =o. ‘The stream funetion of plane creeping motion is thus a bipotential (bikarmonic) function, In the remaining sections of this chapter we propose to discuss three examples of erceping motion: |. Parallel flow past a sphere; 2. The hydrosynainie theory of lubrication; 3. The Hele-Shaw flow. b. Parallel flow past a sphere ‘The oldest known solution for a creeping motion was given by (. G. Stokes who investigated the case of parallel flow pasta sphere [17]. We shail now describe tho result of his calenlations without going into the mathematical details of the theory. We shall base our description on that given by 1. Prandtl [12]. The solutic of eqns. (6.3) and (64) for the caso of a sphere of rading R, the centre of which eo. incides with the origin, and whieh is placed in a parallel stream of uniform velocity Uso, Fig, 6.1, along the 2-axis can be represented by the following equations for the pressure and velocity components: woul (a) 1 Be (6.7) 4 VI. Very slow motion where r? = 2? + y® 4-2? has boen introduced for the sake of brevity. Tt is easy to vorify that these expressions satisfy equs. (6.3) and (6.4) and that the velocity va- niahes at all points on the surface of the sphere. The pressure on the surface becomes 3 P Poo = — Ht fa Von (6.74) The maxinnin and minimum of pressure occurs at points P, and P,, respectively, their values being 3 pve (6.7) 12 — Poo 2° Rk ‘The pressure distribution along a meridian of the sphere ag well as along the axis of ahseissac, 2, is shown in Mig, 6.1. ‘The shearing-stress distribution over the sphere can also be caleulated from the above formulas. 1b is found that the shearing stress has its largest value at point A where t = § fe Uoo/ R and is equal to the pressure rise at P, or presenre decrease at 1s, Integrating the pressure distribution and the shearing stress over the surface of the sphere we obtain the total drag Da Grp RU. (6.8) 6.1. Pressure distribution around a sphere in parallel uniform flow This is the very well known Stokes equation for the drag of a sphere, It can be shown that one third of the drag is due to the pressure distrilmtion and that the remaining twa thirds are due to the existerice of shear, {tis further remarkable that. the drag ie proportional to the first, power of velocity. Ufa drag coefficient is farmed hy referring the drag to the dynumie head } g Uos?,and the frontal area, as is done in the case af higher Reynalds munhers, ar if we put D— Cy RB (YQ Un"), (6.9) thei 2 Cp=7hs R= Veod (6.10) b. Parallel flow past a sphere 5 A comparison between Stokes’s equation and experiment was given in Fig. 1.5 from which it is secn that, is applics only to cases when R < |. The pattern of streamlines in front of and behind the sphere nist be the samo, as hy reversing the direction of free flow, i. e., by changing the sign af velocity components in eqns. (6.3) and (6.4) tho system js transformed into itself. The stroamlines itr viscous flow past. a sphero are shown in Fig, 6.2. They were drawn as they would appear to an observer in front of whom the sphere is dragged with a constant velocity Uso. ‘The sketch contains also velocity profiles at several cross-scelions. It is acen that the sphere drags with it a very wide layer of fluid which extends over abaut ane diameter on hoth sides, At very high Reynolds numbers this boundary layer heeomes very thin. LA ym Ml Bz Fig. 6.2, Streamlines and velocity distri Fig. 6.3. Streamlines in the flow bution in Stokes’ solution for a sphere pasta sphere from Oscen’s so parallel flow tiow Oscen’s improvement: An improvement of Slokes's sohition was yiven by ©. W. Oscen [11], who took the inertia terms in the Navier-Stokes equations partly into acconnt. Te assumed that the velocity components can be represented as the sunt of a constant and a perlnrbation term. ‘hits uaUeg te; va; wow, (6.11) where wn’ and w’ are the perturbation terms, and as such, small with respect to the free stream velocity Uso. It is to le noted, however, that this is not true in the immediate neighbourhood of the sphere. With the assumption (6.11) the inertia torms in the Navier-Stokes eqns. (3.32) are decomposed in two groups, ¢. £.1 au’ ev" au aot co fey Van Sevens and WS we U. 116 VI. Very slow motion ‘The second group is neglected as it is small of the sccond order compared with the first group. ‘Thus we obtain the following equations of motion from the Navier- Stokes equations: ul ’ 2 Veo Se + P= p tw a | Op 27 2 Veo Fe + jy =H VV aoe (6.12) Bw" e . 0 Veo ay x =p Vu au’, av’ ae ey + ee 0. The boundary conditions are the same as for the Navier-Stokes equations, but. the Oseen equations are linear as was tle case with the Stokes equations. The pattern of streaniines is now no longer the same int front of and behind the sphere, This can be recognized if reference is made to eqns. (6.12), becanse if we change the sign of the velocities and of the pressure, the equations do not. trane- form into themeclves, whereas the Stokes eqnatious (6.3) did, The streamlines of the Osecn equations are plotted in Fig. 6.3, and the observer is again assumed to he at rest with respect to the flow at a large «distance from the sphere; i ned that the spltere is dragged with a constant velocity Uso. The flow in front of the sphere is very similar to that given by Stokes, but behind the sphere the streamlines are closer together which means that the velocity is larger than in the former case. Vurtherinore, behind the sphere some particles follow its motion as is, in fact, observed experimentally at large Reynolds mmnbers. ‘The improved expression for the drag coefficient. now becomes U-. op = 4 (14 BR): Rave, (6.13) Experimental results show, Fig. 1.5, curve (2), that Oscen’s equation is applicable up la R= approximately. ¢. The hydrodynamic theory of lubrication The phenomena which take place in oil Inbricated bearings afford anothor example of flow i: whieh viscous forces are predominant. From the practical point of view these phenomena are very important. At high velocities the clearance between two machine elements which are in relative motion (¢. g. jonrnal and heariug) ia filled by au oil stream in which exttemely large pressure differcisces may be created, Asa consequence, the revolving jonrual is lifted somewhat: by the oil film and metallic contact between the moving parts is prevented, ‘The esscutial feattres of this type of motion can he understood on the example of a slide block or slipper moviug ov a plane gnide snrface, Fig. 6.4, it. being important that. they are iuclined ata smallangle, 5, to each other, Weshallassnne that. the sliding suvfaces ¢. The hydrodynamic Ueory of lubrication 7 are very large ina transverse direction with respect. to the motion so that. the problem js one in two dimensionst. In order to obtain a steady-state problem let, us assume that the block is at rest. and that the plane gnide is forced to move with a constant. velocity U with respect to it. The a-axis is assumed in the direction of motion, and the y-axis is at, right angles to the plane of the gnide, The height. h(a) of the wedge between the block and the guide it asained to be very small compared with the length L of the block. Vig. 0.4, Lubrication in a bearing: a) Flow in wodlge hetwoon slide block and plane guide surface; b) Pressure distriltion over block, alt = V7 Tits motion is a more general example of that considered in Se the motion between two parallel flat walls with a pressure gradient. ‘The essen difference consists in the fact that here the two walls arc inclined at an an cach other. For this reason the convective acceleration 1 du/dz ia evidently di from zero. An estimation of the viscous and inertia forers shows immetliately that, in spite of that, in all eases of practical importance, the viscons foreos are predo- minant, ‘The largest viscous term in the equation of motion for the x-direction is equal to pe @ufay?. Hence we can make the following estimate: Inertia force _gudujdz _ p U/l _ g Ut, (4): Viscous force poufey? pt ~ pe t ‘The inertia forces can be neglected with respect to the viscous forees if the reduced Reynolds number : pre Ut (+) él, (6.14) » or, by way of numerical example: U = 40 ft/sec; 4in 4x 10-4 ft%/sce; hk = 0-008 in. 333 ft ” This leads to a value of the Reynolds number referred to the length of the block of U Yr 5,000, whereas the reduced Reynolds number R* = 0-1. + Tho two-dintensional theory was first formated by O. Reyttolds, cf. Phil. ‘Trans.Roy. Soe. (1886), Pt. 1, see also Ostwalds Klassiker No, 218, p. 39. 118 VI. Very slow motion The differential equations of creeping motion, equs. (6.3), can be further simplified for the case under consideration. The equation for the y-direction can be omitted altogether because the component v ia very sinall with respect to u, Further, in the equation for the x-direction %x/8x2 can he neglected with respect to d/dy2, be- cause the formor is smaller that the latter by a factor of the order (h/1)?. he pressure distribution ust satisfy the condition that p == 7% at both ends of the slipper. Compared with the caso of flow hotweeit parallel sliding walls, the pressure graclient, in the direction of motion, Op/2x, is to longer constant, but the very small pressure gradient i in the y-direction cart be neglected, With these simplifications the differenti equations (6.3) reduce to e (6.15) and the equation of continuity in differential form can be repluced by the condition that the volume of flow in every section must be constant: Mz) Q= [ «dy = const , (6.16) 6 The honndary conditions are: y= 0: =U: r=0! p=p (6.17) yok: u=0; xrali p=p. ‘The solution of equ. (6.15) which satisfies the boundary conditions (6.17) is similar to eqn. (5.5), namely uno (i) er y (-4) (6.18) where p’ = dpjdax denotes the pressure gradient, whiclt must be determined in such a way aa to antisfy the continuity equation (6.16), and the boundary conditions for pressure, Inserting (6.18) into (6.16) we first obtain hep’ Oe ign? or, solving for p': p= 2p (om 7 (6.19) Hence by integration P(t) = I +O" uf % a (6.20) 4 Inserting the condition p == py at x — 7 we obtain the vale wife dy, /f de (6.21) ©. ‘Tho hydrodynamic theory of lubrication 119 ‘Thus tlre mass flow is known when the shape of the wedge is given as the function h (), Liqn. (6.19) gives the pressure gradient, and eqn, (6,20) gives the pressure distribution over the slipper. ‘The quantities z z by (a) = f dene amd —by(2) = f dfn, (6.22) 6 é Which appear in eqn, (6.20) depond only on the geometrical ahapo of the gap helaveen the slider and the plane, ‘Their ratio (x) = b, (x)/b, (x) (6.23) which has the dimension of a lengél plays an important part in the theory of lubrication; its vahte for the whole channel, = (hy 1 1 f daphne f deaih) , (6.24) é é is sometimes called the characteristic thickness, With ita aid, the equation of con- tinuity (6.21) can be contracted lo Q=1,UN, (6.25) from which its physical interpretation is evident, The pressure can now be written P(x) = Py + Oe U by (x) — 12 4 Q by (zx) , (6.26) and the pressure gradient becomes: OnU ul a4 (1 — 4), (6.27) Which shows that the pressure has & maxima or a miniinam at a place where the channel thickness is equal to its charaetcristic value, A= JI, Often it is desirable to maintain a positive excess of pressure p — Pp, and the preceding equation can be used to derive the condition for it, Assunting that: P — Py =O atx =O and that the thickness J is placed ale = ay, we nut have h(x) > TI for 0 0 (x) 1a ituplyinng r \ (6.28) h(a) 400 turbulent flow. When the flow becomes unstable, the torque acting on the rotating aylinder increnses steeply, because the kinetic energy stored in the secondary flow mnst be compensated by work, The same flow phonomena, generally speaking, occur when the bearing is loaded and the gap width varios cirenmferentially, but. the details of the flow berome more complex. Attentpts Itave been inade to calouiate the turbnlent low in a gap of a bearing witit the nid of Prandtl'a mixing length (Chap. XIX, oqn. (19.7)]. Tho set of these problems lina attracted a wide circle of investigators, such as D. IF, Wilcock [19]. V. N. Constantinesen (2, 3, 4]. B.A. Saibel and N. A. Macken [14, 15] have written two genoral accounts that contaitt nmmorons literature references, d. The IIcle-Shaw flow Another ranarkable solution of the three-dimensioal equations of erecping motion, equs. (6.3) and (6.4), can be obtained for the case of flow hetaveen two parallel flat walls separated by a small distance 2h. If a cylindrical body of arbitrary cross-section is inserted between the two plates at right angles so that it completely fills the space between them, the resulting pattern of streamlines is identical with that. in potential flow about the same shape. H. S. Hele-Shaw [7] used this method to obtain experimental patterns of strcamlincs in potential flow abont arbitrary bodies. It is casy to prove that the solution for creeping motion from equs. (6.3) and (6.4) possesses the saine streamlines as the correaponding poter flow. We seloct a system of coordinates with its origin in the coutre between the two: plates, and make the z, y-plane parallel to the plates, the z-axis het perpendicular to them. The body is assinned to be placed ina stream of velocity Us, parallel to tho a-axis. Ata large distance from the body the velocity distvilmtion is paruholic, as in the motion in a rectangular chaunel which was considered in Section V.1. Hence ; va(t—g).» A solution of eqns. (6.3) and (6.4) ean be wrilten as: X= 00: u 0, = tg 9) (1 —£) + v= a(x, 9) ( | : (6.39) 2 2 pais J[utenaz=— Me Posen dy | we vilmtion where to(%,¥), %o(%Yy) and po(x,y) denote the velocity aml pressure di 124 VL. Very slow motion of the two-dimensional potential flow past the given hody, This zo, v9 and pg satisly the equations uy dup __ sd am "0 Ge 1% ay =o be au au dL am Me ge bey (6,40, b, e) Bug, a0 az + ay = 9° First, we notice at once from the solution (6.39) that the eqnation of continnity and the equation of motion in the z-direction are satisfied. The fact that the equations of motion in the x- and y-«irections are also aatisfied follows from the potential character of vg and 4. The functions 19 and vp satisfy the condition of irrotationality Gugl?y — dv_lax = 0, so that the potential equations V2 tq = 0 and V2 = 0, where 7? = 2/ax? + @%/ay2, are satished. ‘The first two equations (6.3) reduce Lo Apjdx = je 0u/dz* and Opfay = p22*v/dz", they are, however, satisfied, aa scen from cyns. (6.39). ‘Thus oqns. (6.39) represent asolution of the equations for creeping motion. On the other hand the flow represented hy ccns. (6.39) has the samo streamlines as potential flow abont the body, and the alreamliues for all parallel layers 2 = const aro congruent. 'The condition of no slip at the plates z = +h is seen to be satisfied by eqn. (6.39), but the condition of no slip at the snrface of the body is not satisfied. The ratio of inertia to viscous forces in Hele-Shaw motion, just as in the case of the motion of Inbricating oil, is given by the reduced Reynolls number Uk 7 nea Vol (A) ea, ¥ teristic linear dimension of the body in the 2, y-plane. ms becoine considerable and the motion deviates where L denotes a chara If R* exceeds unity the inertia t from the simple solution (6.39). The solution given hy eqn. (6.39) can be improved in the samc manner as Stokes’s solution for a sphere or the solution for very slow flow. The inertia terms are latel from the first approximation and introduced into the equations as external forces, und an improved solution results. This was carried out by F. Riegels [13] for the case of Hele-Shaw flow past a circular cylinder. For R* > 1 the streamlines in the various layers parallel to the walls cease to be congriont. The slow particles near the two plates are deflected more by the presence of the body than the faster particles near the centre. ‘This causes the streamlines to appear somewhat, blurred and the phenomenon is inore pronounced at the rear of Uke body than in front, of it, Fig. 6.6. Solutions in the case of eroeping motion are inherently restricted to very small Reynolla numbers. In principle it is possible to extend the fick! of application Reforences 128 to larger Reynolds numbers by successive approximation, as mentioned previously. Mowever, in all cases the calculations become so complicated that it is not practicable to carry out more than one step in the approximation. For this reason it is not possible to reach the region of moderate Reynolds numbers from this direction. To all intents and purposes the region of moderate Reynolds numbers in wh tho inertin and viscons forces aro of comparable magnitude throughout the ti of flow has not. been extensively investigated by analytic means. It, is, therefore, the more useful to have the possibility of integrating the Navier-Stokes equation for the other limiting case of very large Reynolds nimbors. , Thus we are led to the bomndary-layer theory which will form the subject of the following chapters. Fig. 6.6. Hele-Shaw flow past circular cylinder at R* — 4, after Ricgels [13] References: {1] Bauer, K.: Einfluss der endlichen Breite des Cleitlagers auf Tragfthigkeit und Reibung. Forsohg. ing.-Wes. /4, 48—62 (1943). [2] Constantinescu, V.N.: Analysis of boarings operating in turbulent regime. ‘Trans. ASME, Series D, J. Basic Eng. 84, 139—151 (1962). {3] Constantinescu, V.N.: On the influcnce of inertia forces in turbulent and Iaminar self. acting films. Trans. ASME, Series F, J. Lubrication Technology 92, 473--481 (1970), [4] Constantinescu, V.N.: On gas Iubrication in turbulent regime. Trans. ASME, Series D, J. Basic Eng. 86, 475-482 (1964). (5) Fréasel, W.: Reibungawiderstand und Tragkraft cines Gleitschulies endlicher Breite. Forschg. Ing.-Wes. 13, 65--75 (1942). {6] Giimbel, L., and Everling, [.: Reibung und Schmierung im Maschinenban, Berlin, 1925. tH Hole-Shaw, H.S.: Investigation of the nature of surface resistance of water and of stroa motion der certain experimental condi ons. ‘Traus. Inst. Nav. Arch. X7, 2 (1898); seo also Nature 58, 34 (1898) aud Proc. Roy. Inst. 16, 49 (1809). [8] Kahlert, W.: Der Hintluss der Tra elt ie ‘bel'der hydrodynamisehien Schnuiermittel- theorio. Diss. Braunschweig 1947; Ing.-Arch, 76, 321—342 (1948). {9] Michell, A.G.M. Z. Math. t. Phys. 52, 8. 123 (1905); see also Ostwald’s KInasiker No. 218, [10] Nabine, F.: Boitriige zur tiydrodynamischen Theorie dor Lagerreibung, Ing.-Arch. 17, 191--209 (1940). [11] Oseen, C.W.: Uber die Stokes'sche Forme! und iiber eine verwaudte Aufgabe in der Hydro- dynamik. Ark. £ Math, Astron. och Fys 6, No. 29 (1910). [12] Prandtl, L.: The mechanics of viscons fluids. lu W.F. Durand: Acrodynamic Theory /1/, 34—208 (19 [13] Riegels, F.: Zar Kritik des Hele-Shaw-Versuches. Diss, Gottingen 1938; ZAMM 18, 95— 106 (1938), (14] Saibel, B.A, and Mackon, N.Avt The fluid mee! id Mech. (M. Van Dyke, ed.) 5, 185—212 (1973), [15] Saibel, E.A., and Macken, N.A.: Non-lamisar behavior in bearings. C literature, Trans, ASME, Series F, J. Lubrication Technology 96, 174 ies of lubrication. Aunual Review of ‘al review of the -181 (1974), 126 VI. Very slow motion 116] Sommerfeld. A. Zar hydrodynamischon ‘Theorlo der Selimiorniittelroibung. Z. Math. uw. Physik 50, 97 (1904); also Qstwald's Klassiker No. 218, p. 108, aud: Zur Theorie der Schitior- niittelreibung. Z. ‘Techn. Phys. 2, 68 (1921); also Ostwald's Klassiker No. 218, p. 181. [17] Stokes, GG. On the elfect: 0 ° om the motion of pendalums. ‘Trans, Jamnbr. 08 (1851) or Coll. Papers 171, 55. (18] Taylor, G.1.: Stabilily of » viscous liquid contained betw'con two rotating cylinders. PI ‘ 5 289 — 203 (1923). (19] Wilcock, D. I. Turbulence in high-speed journal benrings. Trans, ASME 72, 826 (1950). [20] Vogolpolil, G.: Beitrige zur Keuntuis der (loitlagerreibung. VDI-Forachinigsheft 386 (1937). [21] Yogelpohl, G.: Abnlichkeitabeziehungen der Gleitligerreibung und untere Reibungagrenze. 2. VDL 91, 379 (1949). {22] Vogelpohl, @.: Betriebssichere Gleitlager. Bercchunugsverfalren fir Konstruktion und Betrieb. Vol. U, Springer-Verlag. 2nd, ed, Berlin, 1967. Part B. Laminar boundary layers CHAPTER VII Boundary-layer equations for two-di sional incompressible flow; boundary layer on a plate a, Derivation of boundary-layer equations for two-dimensional flow We now proceed to examine the secoud limiting ease, namely that of very small viscosity or very large Reyuolds number. An inportaut contribution to the seience tion was made by L. Prandtl [21] in 1904 when he clarified the essential nolds mambers and showed how the ier-Stokes equations could be simplified to yickl approximate solutions for We shall explain these simplifications with the aid of av argimiont which. preserves the physical picture of the phevoinenon, and it will be recalled that in the bulk of the fluid inertia forees predominate, the influence of viseons forces being vanishingly small, Vig. 7.1. Boundary-layer How aloug a wall eusiontl flow of a thie Vor the sake of simplicity we shall consider two-di wilh very amall viscosity: about 2 cylindvieal body of sleuler eros: With the exception of the immediate neighbourhood of the sur! are of the order of the free-strean velocity, V, and the patter “ distribution deviate only slightly from those i ionless (potential) ver, detailed vestigations reveal that, unlike in potential flow, the fluid does not slide over the wall, but adberes to it. The transition from zere velocity at the wall to the fall wnagnitade at some distance from it takes phice in a very in 128 VLE. Boundary-lnyer equations for two-dimensional flow; boundary layer on a plate layer, the so-called boundary layer. In this manner there are two regions to consider, even if the division between them is not very sharp: 1. A very thin layer in the immediate neighbourhood of the body in which the velocity gradient. normal to the wall, du/ay, is very large (boundary layer). In this region the very small viscosity jt of the fluid exerts an essential influenc in so far as the shearing stress r= je(Mufdy) may assume large values. 2. In the remaining region no sich kuge velocity gradients oecur and the inflien of viscosity is uniwportant. In this region the flow is frictionless aul potential. In general it is possible to state that, the thickness of the boundary Iayer in- creasea with viscosity, or, more generally, that, it decreases as the Reynolds vumber increases. It was secp from several exact solutions of the Navier-Stokes equations preseuted iv Chap. V that the boundary-layer thickness is proportioval to the square root. of kinematic viscosity br When making the simplifieations to be introdnecd into the Navier-Stokes equations it is assumed that this thickness is very small compared with a still unspecified linear dimension, 1, of the body: o 0) can he easily inferred from 2 consideration of the relation betwecn the pressure gradient. dp/da and the velocity distribution u(y) with the aid of the boundary-layer + The velocity profile at. Lhe point of separation is sect to lave a porpendioular tangent at the wall. ‘The velocity profiles downstroam from the point of separation will show regions of rovers: ed flow near the wall, Mig, 7.2c. c. A retnark on the integration of the boundary-layer equations 133, v= 0 we have ab au\ _dp a er a and, firther, after differentiation with respect lo y: (7.16) In the inunediate neighbourhood of the wall the curvature of the velocity profile depends only on the pressure gradient, and the eurvature of the velocity profile at the wall changes its sigit with the pressure gradicnt. For flow with decreasing pressure (accelerated flow, dp/dz < 0) we have from eqn, (7.15) that (22/2y%)wan <0 and, therefore, %u/dy? <0 over the whole width of the boundary layer, Vig. 7.3. Tu the region of pressure increase (decelerated flow, dp/de > 0) we find (2?u/2y2) > 0. Sinoe, however, in any case d®u/dy? <0 at a large distance from the wall, there inust exist a point for which 22x/dy? = 0. This is a point of inflexiont of the velocity profile in the bowndary layer, Pig. 7.4. y y y i S A au tu eSB ES na Fig. 7.3. Velocity distribution in a boundary Fig. 7.4. Velocity distribution in a boundary layer with pressure decrease layer witht pressure increase; Pl = point of infle: It follows that in the regiou of retarded potential flow the velocity profile it the bonudary layer always displays 2 point of inflexion, Since the velocity profile at the point of separation and with a zero tangent mist. bave n point of inflexion, it follows that separation can only o when tho potential flow is retarded c. A remark on the integration of the boundary-layer equations Int order to integrate the homtdary-layer equations, whether in Ute non-steudy case, eqns. (7.7) and (7.8), o in the steady case, eqns. (7.10) and (7.11), it is often convertiont to introduce a stream function p(x, y,¢) defined by ap By? (7.17) + The oxistence of a point of inflexion in the velocity profile in the boundary layer is important for its stability (transition fron laminar to turbulent flow), soe Chap. XVI. 134 VIE. Boundary-layer cquations for two-dimensional floy dary layer on a plate 60 that. the equation of continuity is thereby satisfied. Introducing this assumption into equ. (7.7) we have an dy at ep ay dy ay Lap Op dy dzay ~ dz dy? o ax 1’ ays’ (7.18) which is a partial differential equation of the third order for the stream function. The boundary, conditions require the absenes of slip at the woll, or Ay/ay = dy/esx ot the wall. Farther, the initint condition at ¢ = 0 prearrihes the velocity distribntion w= dy/éy over the whole region. Tf thia equation for the styeam finotion is eomparod with the complete Navier-Stokes equations (4-10), it is seen that, the boundary-layer assniplions have reduced the order of the equation from four to threo. d. Skin friction When the boundary-layer equations are integrated, the velocily distribution can be deduced, and the position of the point of separation can be determined. This, in turn, permits us to calculate the viacons drag (skin friction) around the surface by a sinple process of integrating the shearing stress at the wall over the surface of the hody. ‘The shearing stress at the wall ix (an TOE \ By) yoo” al flow becomes ‘The viscous drag for tho case of two-dimension ! b f 1» 008 pds , (7.19) where b denotes the height of the cylindrical body; 6 is the augle between the tangent to the surface and the free-streain velocity U,., and sis the coordinate measured along the surface, Fig. 7.5, ‘The process of inttesration is to he performed Fig. 7.5. Mustrating Une ealeutation of skin friction over the whole surface, from the stagnation point at the leading edge to the trailing edge, assmming that there is no xeparation. Since cos ¢ ds == dx, where 2 is mncagured parallel to the free-stream velocity. we can also’ write ' 1 aul Dy, nf (3) jae, (7.20) so and the integration, as before, is to he extended over the whole wetted surface from the lending to the trailing edge, In order to calculate the skin friction it. is necessary ©. The boundary layer along a flat plate 135 to know the velocity gradient. at the wall, whieh can be achieved only throngh the integration of the differential equations of the boundary layer. If separation occurs before tho trailing edg valid only as far as the point of separation. Furthermore, if the lam ‘y layer transforms into a turbulent one, eqn. (7.20) applies only as far as the point of transition, Beltind the point of transit there is turbulent friction, to he discnssed in Chap. X X11, If separation exists, the pressure (731) with (7(0) =a = ‘332 from Table 7.1. Henee the dimensionlesa shearing stress becomes: (7,32) fy D=apbUcol/ ie =2ab Delft @t Us» rio and for a plate wetted on both sicdes¢ 2D =4abUm Vl wet = 1.328 b VU pol. (7.38) Th fs voma whereas ble that. the skin friction is proportional to the power § off velocity v ereeping motion there was proportionality to the finst. power of velo Fhrther, the drag inercases with the square root of the length of the plate, ‘This can be interpreted as showing that the downstream portions of the plato contrfbute proportionately less to the total drag than the portions near the leading edge, c. The boundary layer along a flat plate 139 Table 7.1. The function {() for the boundary layer along a {Int plate at zero Incidence, after L, Howarth [16] es at 7 "OV oe J T= ae f 6 0 0 033206 O2 0-00664 006641 0:33199 o4 002656 013277 0-33147 . 0-6 0-06974 0-19894 033008 O8 0-10611 0-26471 0°32739 10 0-16557 0-32979 0°32301 12 023795 0°39378 031659 14 032298 0°45627 030787 16 0-42032 0°51676 0°29667 18 0°52052 057477 0°28293 20 065003 0-62977 026675. 22 078120 068132 0-24835 24 092230 0-72899 0-22809 26 107252 0°77246 0°20646 28 123099 081152 018401 30 139682 084605 016136 32 156911 087609 013913 a4 1-74696 0-90177 0-11788 36 1-92954 092333 009809 38 2-11605 0-94112 008013 40 2-30576 0-95552 006424 42 2-49806 096696 005052 44 2-69238 097587 003897 46 288826 098269 0-02948 48 3-08534 0-98779 0°02187 50 3-28329 0-99155 001591 52 348189 0-99425 001134 4 3.08004 | 0-00616 | 0-00793 56 3°88031 099748 000643. 58 407990 0-99838 000365 60 427964 0-99898 000240 62 447948 0-:99937 000155 6-4 4-67938 0-99961 000098 66 487931 099977 0-00061 68 6-07928 0-99987 000037 70 27926 0-99992 0-00022 72 5°47925 0-99996 0-00013 14 667924 0-99998 0-00007 76 87924 099999 000004 78 607923 1.00000, 0-00002 80 627923 1.00000 000001 82 6-47923 100000 000001 84 6-67923 1-00000 000000 86 6-87923 1-00000, 000000 88 707923 1-00000 000000 140 VIF. Boundary-layer equations for two-dimensional flow; boundary layer ont a plate because they lic in the region where the boundary layer is thicker and where, eonse- ly, the shearing stress at the wall is smaller, Introducing, as usual, a dimen- where A = 21 denotes the wotted surface area, we obtain from eqn. (7.33) the formula: (7.34) Here R, = Uso l/y denotes the Reynolds number formed with the length of the plate and the free-stiream velocity. This law of friction on a plate first deducod by IT. Blasius, is valid only in the region of laminar flow, i. ¢. for Ry = Uso fy <5 x 10% to 10*. It is represented in Fig. 21.2 as curve (1). In the region of turbulent, motion, R, > 108, the drag becomes considerably greater than timt given in equ. (7.34). Boundary-layer thickness: It; is impossible to indicate a boundary-layer thickness in an nnambignous way, beoause the influence of viscosity in the boundary layer decreases asymptotically outwards. The parallel component, 1, tends asymptotically lo the value Uo. of he potential flow (the function f’ (7) tends asymptotically to 1). If it is desired to define the bonndary-layer thickness as that distanee for which 0-99 Uo, then, as seen from Table 7.1, 7 5-0. Hence the borndary-layer kness, as defined here, becomes swoo te. (7.36) A physically incaningful measure for the boundary layer thiel ment thickness 54, which was already introduced in cqn. (2.6), Fig. 2.3. ‘The dis- placement, thicknoss is that distance by which the external potential field of flow ia displaced outwards as a consoquence of the decrease in velocity in the boundary is f (Toa —te) dy, 70 is the displace- layer. The decrease in volume flow due to the influence of fri so that for 6, we have the definition Uonby = f (Uno) dy, or “° =f (i (7.36) fl With 1/Ton from eqn, (7,26) we obtain n= Vie J —rentan = 32 bn fond. e. The boundary layer along a flat. plate 141 where 7, denotes a point outside the boundary layer. Using the value f(7) from Table 7.1 we obtain m, -- f(7,) = 1-7208 and itence 5, =1-7208 Viz (displacement thickness), (7.37) ‘The distance y = d, is shown in Fig. 7.7. ‘his is the distance by which the stream- lines of the external potential flow are displaced owing to the effect of friction near the wall. The boundary-layer thickness, 3, given in eqn. (7.36), over which the potential velocity is attained to within | per cont. is, in round figures, three times larger than the displacement thickness. We may at this point later. ‘The loss of momentum luate the momentum thickness 2 which will be used the boundary layer, as compared with potential flow, is given by Qf «(Uoo—u) dy, 80 that a new thickness can be defined by é 0 co? bg = 9 f u(Uay — u) dy , yao or 3b, = Se (1 — it) dy. (7.38) =o Numerical evaluation for the plate at zero incidence gives: a= Ve f ronan, ae oss y/ F? (momentum thickness). (7.39) or 6, It is necessary to remark here that near the leading edge of the plate the bound- ary-laycr theory ccases to apply, since there the assumption | a%u/ax? | < | %u/ay? | is not satisficd. The boundary-layer theory applics only from o certain value of the Reynolds number R = Us. 2/v onwards, The relationship car the leading edge can only be found from the full Navier-Stokes equations because it involves a singularity at the leading edge itself. An attempt to carry out such @ calculation was made by G. F. Carrier and C. C. Lin [6] as well as by B. A. Boley and M. B, Friedman [3]. Experimental investigations: Measurcmonta to test the theory given on the preceding pages were carried out first by J. M. Burgers [4] and B. G. van dor Hegge Zijnen (15), and subsequently by M. Hansen [14]. Particularly carcful and com- prehensive measurements were reported later by J. Nikuradse (20). It was found that the formation of the boundary layer is greatly influenced by the shape of the leading edge as well as by the very small pressure gradient which may exist in the 142 VIL. Boundary-layer equations for two-dimensional flow; houndary layor on a plate external flow. J. Nikuradse introduced careful corrections for these possible effects, when he carried out his measurements on a plate ina stroan of air. ‘The velocity distribution in the laminar boundary tayer has been platted from Nikuradse’s measurements in Fig, 7.9 for soveral distances fram the leading odge, ‘The similarity 10 08 10 20 10 40 50 distrihution in the laminar boundary layer on a flat plate at zero a8 measured by Nikuradse [20] icidenee, of the velocity profiles at varions distances x from the leading edge predicted by the theory is confirmed by these measuraments, The shape of tho velovity profile agrees equally well with that. calenlated with the aid of the theory. ‘The relation between the dimensionless bonndary-lnyer thickness 6 Vf Uso/y x and the Reynolds number formed with the enrrent length, «, was already plotted in Fig. 2.19, ‘This dimensionless thickness rei long as the baundary layer is laminar, and namerical vahic 4 given in equ, (7.35). Ab large Reynolds numbers Uo x/¥ the ho to bé laminar and tran motion takes place. This fact can he recagnized in Fig. 2.19 hy notieizg the inarked increase in the thickness of the bomdary layer as the distance from the leading cdgo is inereased, According to the measurements performed hy B.C. van der Hogge Zijnen anc M. Hansen transition from laminar to turbnlent flow takes place at Hos /y = 300,000. This corresponds to a value of the Reynolds number referred lo the displacement thickness, Uo, 5,/» = 950. More recent measurements, to be jon ta turhnlent e. The houndary layor along a flat plate 143 discussed in Chap, XVI, have demonstrated that number can become value of this ‘critienl” Reynolds iderably larger in an air stream whieh is made very free from distnrbanee. Tn this way it is possible ta reach valnes of up ta ahont um “3x 108, Fig. 7.10. Local eoefficient of skin fi ph jos on a flat Je al zero incidenoo in . conipressible flow, deter- .0005 ad from direet. mensure- ment of shearing stresa by 0003 and Dhawan Indirect skin friction measurement from velocity profile © Direct skin friction measurement, x = 28.6.em ‘Theory: Inmminar from eqn. (7-12); Uarhotent front eqn, (21.12) The laminar law of friction on a flat plate was also subjected to careful ex- rental verification. The local shearing stress at the wall can be determined indireetly from the slope of the velocity profile at the wall together with eqn. (7.31). In recent times H.W. Liepmann and §. Dhawan [18] measured the shearing stress directly from the foreo acting on a small portion of the plate which was arranged so that it could move slightly with reapect to the main plate. The results of their very careful measurements are seen reproduced in Fig. 7.10, which shows a plot of ‘the local coefficient of skin friction o/’ = to/f @ Uo, against the Reynolds numbor R, = Uso x/v. In the range of R, = 2 x 10% to 6 x 10% both laminar and turbulent flows are possible. It can be scen that direct and indirect: moasurements are in excellent agrecinent with cach other, Mensuremonte in tho laminar range give iking confirmation of Blasius’s equ. (7.32) from whieh ¢’ = O-604/VR,. In the turbulent range. re is also good agreement with Prandtl’s theoretien! formula whieh will be deduced in Chap. XXI, eqn. (21.12). Tho camplete agreement between theoretical and experimental results which exists for the velocity distribution and the shearing stress in a laminar bomdary layer on a flat, plate at zero incidence that. has been brought. into evidence in Figs, 7.9 and 7.10 for the range Rz > 105 nnequivocally demonstrates the validity of the boundary-layer approximations from the physical point of view. In spite of this, 144 VET, Boundary tayer equations for two-dimensional flow; boundary layer on a plate certain inathematicians have expended much effort to create a “inathematical proof” for the validity of these simplifications; in this connexion consult the work of H. Schmidt and K. Schroeder (24). f. Boundary layer of higher order} ‘The boundary-layor equations have been obtained in Sec. VI a of this chapter by a process of ostimating orders of magnitude of individual terms in the complete equations of motion, The boundary-layer equations can, however, also be dorived with the aid of a more general theory. In order to obtain asymptotic expansions of the solutions of the Navier-Stokes equations for large Reynolds numbers, it is possible to ostablish a perturbation scheme in wi as 1 ve y= Ro (7.40) is choson as the perturbation parameter. This leads to so-called singular perturbation scheme and results in the separation of the required asymptotic expansion of the solution into an outer expansion (external flow) and an inner expansion (boundary-layer flow). the aid of the method of matched asymptotic expansions it thus becomes possible to derive an asympto- tic expansion of the complete solution. ex The first term of such an asymptotic expansion is procisely Ute solution of the boundary. layer equations. Moreover, the continuation of the perturbation ealeulation allows us to compute further terms of the expansion and s0 to extond the classical theory of boundary layera duc to PrandU. We thus create a boundary-layor theory of higher order. "He sovoncl terme of tho ox. pansion are of particular practical importance because we can look upon them as corrections to the classical theory wltich represent boundary-layer effects of second order. Extensive presentations of boundary-layer theory of higher order were published by M. Van Dyke [9], K. Gerston [J0], and K. Gersten and J. F. (Grose {12}. In addition, reference [8] con- tains a detailed exposition of the method of matched asymptotic expansions. The basic ideas of this method can be traced to L. Prandtl; they have been made plausible with reference to a simple mathematical example in Soe. LVf. In what follows, we’give a brief deseription of the theory of asymptotic solutions for large Reynolds numbers as it applies to a two-dinensional, incompressible flow. The inain purpose of this argument is to find an extension of Prandtl’s boundary-layer theory and to derive the bound- ary-layer equations of higher order. Details of the derivations can be found in the treatise of M: Van Dyke [7]. ‘The starting point is constituted by the Navier-Stokes equations written with reference to ‘a curvilinear, rectangular system pf coordinates in See. I1Jg, Fig. 3.9. All longths are measyred in units of a convenient length Ro, for example the radius of curvature at the stagnation point, Velocities are referred to Us and the overpressures are referred to 9 UZ. The geometrical sliape is doeoribed by the local radius of curvature, R(x), and the dimensionless curvature of the sneface is K(x) = Rof R(x). (7.41) Outer expansions: In order to solve the systein of equations (3,38), we assume the following asymptotic expansion u(a, yy e) = Us(a, y) + & Ura, y) +... v(t, ye) = Vila, y) + & Vola, y) +... (7,42) . P(x, y&) = Pilz, y) + € Paley y) +... Those forms are substituted into eqns. (3.38) and the terms are ordered by the powers of &, In this manner, we obtain a sequence of systems of equations for the first-order solution Ui (x, y)y t Lowe this section to Professor K, Gersten. f. Boundary layer of higher order 145 V(x, 9), Pu(2, y), for the second-order solution Ua(z, ), Va(x, yh. Pala, y), etc. Up to solutions of the second order, terns proportional to e%, that. la the frictional terms in the Navier-Stolc equations, remain unaccounted for. ‘Thus, solutions of first and secénd order correspond to flows or even to potential flows when only fields with a uniform oncoming velocity are stud ‘The solutions of first order entisly the boundary cond y~ 0: Vi(x,0)—0 yoo: URL VE=1. (7.43) ‘The sohition of the potential-low equations Ur(x, y), P(x, y) gives the velocity Us (x, 0) at the wall, and Bernoulli's equation leads to tho wall pressure, Lo: Pil, 0) = 3 ~ = UFO). (7.44) ‘The solutions of second order satisly the boundary conditions Lad yO: Val, 0) = ae (Uae, 0) A1(7)) (7.45) yoo: UR+VE=0, where 51 (x) denotes the displacement thickness defined in an analogous way as that in eqn (7.36); see nlao eqn. (7.51). The solution of the potential equation leads again to the distribution of the parattel velocity components at the wall, U2(z, 0), and to the pressure P2(x, 0) = — Us (x, 0)» Ua(r, 0). (7.46) The resulting solutions do not, generally speaking, satisfy the no-slip condition at the wall and for this reason they are not valid near it; they are given the name “outer solutions” or “outer asymptotic expansions”. Inner expansions: In order to obtain solutions valid near the wall, it is necessary to apply a special procedure, Instead of the distance, y, from the wall, we introduce a new, stretched co- ordinate N = gle. (747) This so-called! inner varinble was so selected as to prevent the disappearance of some of the viscous terms in the equations of first order in the coordinate system x, N. For the solutions near the wall (in the boundary Inyer), we again assume asymptotic ex- pansions, viz, w(x, yy) = ta(z, N) + ¢ur(e, N) +... u(x, ye) == evr(a, N) 1 et va(a, N)A os (7.48) P(x, y, 6) = pr(x, N) + € p2(x, N) + ..~ Substitution into the system of equations (3.38) and ordering according to powers of e, yiolds the following systems of equations, Boundary-layer equations of first order: Ou Ou oz ton au au am | (7.49) “oa *% oN ~~ ae tb oNT’ o = — 2m on’ 146. VI1, Boundary layer equations for two-dimensional flows boundary layer on a plate with the boundary conditions uy = 0,4 = 0, (7.50) N-> co: ay = U(r, 0) ‘These are exactly Prandt!’s bou ver equations, eqns, (7.10) and (7.11) transforined to coordinates x, V. In addition py (x) 0. ‘The solution m1 (2, 4’) allows ns to compute the displacement tr ‘ness 6) defined as ye f(i - ae ah) aw. (7.51) ‘The equations of first order, eqns, (7.49), do not contnin the Reynolds mimber expt follows that ay(r, NV) and ila, ¥) mutat also be, proves that the location of the point of ta Hy. It ndependent of the Reynolds number.“ inar separation is independent of fhe Reynolds nun per. Boundary-layer equations of second order: ér éuz en Cus mS +e tne ex ex én (7.52) K with the boundary conditions N=0: w= Or N > 00; 2 = Ua(x,0) ~ K Ui(2, 0) N, (7.53) pe = Pe(x. 0) 4 K UF (2,0) N, ‘The outer botndary conditions (i. ¢. for N -> 09) of the inner solutions as well as the imer bound- ary conditions of the omtter solitions (c.g. eqn. (7.45) for V'2(x, 0)) follow from the matehing of the inner and onter solutions; see also [7]. ‘The ayatem of equations (7.52), (7.53) for the sccond-order boundary layer 100 docs not contain the Reynolds mmber explicitly. However, it contains solutions of first order and is more extensive than the first-order aystom, but it consists of linear differential equations, For this reason, it is possible, in titrn, to scparate the whole solution into a sun of partial solutions. Tt has become customary to aplit the solution into a enrvatarre term and into a displacement term, Dut we shall not purse thia discussion any further here. Due to the fact that the curvature of the wall is neeounted for in tho second-order theory, there appears a pressure gradiont in the direetion normal to the wall, For this reason, the presaure At the wall becomes different from lint which is impressed on the boundary layer by the outer flow. Inte tos the boundary layer, we obtain the pressure coefficient at the wail in the form. > Spe ~ p(x, 0,6) 2 (7.54) Ps(r0) Le (Pole 0) 1K f [OF r 0) = neler, MAN) 1-0 (2). é ‘The pressiire at the wall exceeds the impressed pressure when the wall ia convex (K > 0). £, Boundary layer of higher order 147 ‘The distribution of the local shearing stress to second ordor is J. ro(2) (2. due r 5 (7.55) 20 OUR woo TON RW yen FO) (7.58) ‘The boundary Inyer of second ordor also reacts on the outer flow. The paper by K. Gerten [11] contains a calculation of the displacement thickness to second order. Example: Flat plate at zero incidence. In the case of an imperincable fiat plate at zero inci. dence, the displacement thickness 51 is calculated with the aid of eqn, (7.37). According to eps. (7.45), the boundary condition for the outor Hos Vo(x, 0) = 0-8604/)'r , (7.56) where the Iength of tho plate has heen chosen a8 n reference. ‘The solution of the two-dbnensionnt potential equation subject to this houndary condition yields Vala, y) ~ — 28804 — (7.57) 08004 7 fra Yat, = \s : where r=Vrry ye. (7.58) ‘The associated streanitines are parabolnc whose foci are at the origin and whose vertices lie on the z-axis, It follows that in this particular ease the velocity Uz(x, 0) at the wall yanislics, aud the solution of the syatem of equations (7.52) and (7.53) ia the trivial solntion, We conclude, therefore, that in the case of the flat plate the second-order correction to skin friction vanishes, Nevertheless, we must not draw the conclusion that the second-order drag coefficient also vanishes. This is due to the fact that the second-order external flow described by eqn. (7.57) contsi- bntes a momentum term. This can be identified by ealonlating the integral of moinentinn over the whole plate when it will be discovered that this contribution is equivalent to an increase in drag, Such calculations have been carried ont by 1. linai (17] who fonnd that the drag coefficient, ofa flat plate is given by gy = 328. 2826 a9 as x a whore 2-326 = 2 x (0-8604), ‘Tho correction (the second term) in can. (7.59) amounts to 5-5% At Ry = 103, docrensing to 2% at. Rr = 108, compared to the first term, ‘The fact that the second term in equ. (7.59) does not represent siin friction is explained by the observation that the singular character of the flow at the leading cclge induces a preesure drag. Presuinably, at the leading edge there arises an infinite overpressure which contribites # finite force in spite of the vanishingly small plate thickness. In this connexion a comparison with the case of the parabola of See. 1X} shontd be made, Striotly apenking, the proceding nnntysi of flow pust a fint plate i restricted tom sani. infinite plnte. In the case of x finite Tength, the shearing stress becomes inodiied mtn certvin Mistance upstream of the trailing edge. However, Prandtl» boundury-layer equations, being parnholic, eannot account for this “trailing-edge offe According to K. Stewartson [26a], it is possible to master such trailing-cdge effects, or, generally speaking, the effects which are expressed as singularities (c. g. lending edge, trailing edge, separation) by Prandtl’ equations, through a generatization of Prandtl’s concept of the boundary Inyer. This is done by the introduction of the iden of “nuultistrncturcd” bormdnry layers of the “triple-deck” concept. For the case of a fat plate, again, K. Stewartson [25a] and A. I, Messiter [18b] find that the akin- ent is givon by 148 VIE. Boundary tayer eqnations for two-dimensional flow; boundary layer on a plate fig. 7.11. Skin-friction eoeffi- cient. of a flat plate of finite tongth at zero incidence (1) Theory after H. Nlaslus, ean.(7.34) 2) Theory after ALF. Menaiter [8b], eqn. G60) A. Theory: after Desbiie (solution of Navier-Stokes edpations) 0 9 W600 000 G00 “00 © Experinenta afure Z, Sanour {301 —=R, cy we E828 | 2-608 (7.60) VR ' (Ry Here, the trailing edge has boon accounted for, but not the displacement effect. The diagram in Fig. 7.11, reproduced from the work of It. E. Metnik and R, Chow (18a), shows that tho values of cy computed with the aid of eqn. (7.60) agree very wall with the results obtained from the complete Navier-Stokes equations as woll aa with thoso of measurements down to Rr = 10. At Rr = 40 eqn. (7.60) loada toc, = 0316 which is lesa than 2% in oxcoss of the oxact vaine ey = 0-311. Section UXj will return to the diseussion of exact solutions of boundary-layer equations of sceond order. References [1] Bairstow, L.: Skin friction, J. Roy. Aero. Soe. 19, 3 (1926). {2] Biasius, H1.: Grenzachichton in Fliissigkeiten mit kleiner Roibung. Z. Math. Phys. 56, 1—37 (1908). Engl. transl, in NACA TM 1256, {3) Boley, B.A., and Friedman, M.B.: On the viscous flow around tho leading edge of a flat plate, JASS 26, 453-454 (1959). [4) Burgers, J.M.: The motion of a finid in the boundary layer along a plane smooth surfnec. Proc. Kirst Intern, Congr. of Appl. Mech., Delft 1924 (C. B. Biozeno and J. M. Burgers, ed.) Delft, 1925, pp. 113--128. {5] Carrier, G.I, and Lin, C.C.: On the nature of thé boundary layer near the leading edge of n flat plate, Quart. Appl. Math. V7, 63—68 (148). {6] Dhawan, S.: Direct mensurements of a ction. NACA Rep. 1121 (1953). {7} Van Dyko, M.: High in boundary layer theory. Part 1: General analysis. JEM 14, 161—177 (1962). Part 2: Application to leading edges. JM 14, 481 —405 (1962). Part 3: Parabola in uniforin stroam. J FM 19, 145—150 (1964). (8) Van Dyke, M.: Perturbation methods in fluid mechauies, Academie Press, New York, 1964. [9] Van Dyke, M.: Highororder boundary layer theory. Annual Review of Finid Mesh. J, 245 292 (1969). References: 149 [10] Gersten, K.: Grenzschichteffok te hoherer Ordnung. Anniversary volume commemorating Professor H. Schlichting’s 65th anniversary (Sept. 30, 1972). Rep. 72/5 Inst. f. Sursmunge- mech. Techn. Univ. at Braunschweig, 29-53 (1972). [11] Gorsten, K.: Die Verdriingungedicke bei Grenzachichten hiherer Orduung, ZAMM 54, 165-171 (1974), [12] Gorsten, K., and Gross, J.F.: Wigher-order houndary layer theory, Phrid Dynamics Trans. netions (1975). [13] Goldstein, S.: Goneerning soine solutions of the boundary layer equations in hydrodynamics. Proc, Cambr. Phil, Soe. 26, | ~30 (1930); see also: Modern developments in fluid dynamics, Vol. 1, 135, Oxford, 1938, [14] Hanson, M.: Dic Geachwindigkeitaverteilung in der Grenzachicht an ciner cingetauchten, Platte. ZAMM 8, 185—199 (1928); NACA 'TM 585 (1930). [15] Van der Hegge-Zijnen, B.G.: Meastirements of the velocity distribution in the boundary layer aloug a plane surface. Thesis, Delft 1924. {16] Howarth, L.: On the solution of the laminar boundary layer equations. Proe. Roy. Soc. London A 164, 547-579 (1038). [17] Imai, 1.: Second approximation to the laminar bonndary layer flow over a flat plate. JAS 24, 150 — 156 (1957). (18) Liepman, H,W., and Dhawan, S.: Direot measuroments of local akin friction in low-speed and high-speed flow. Proc. Firat US Nat. Congr. Appl. Mech. 869 (1951). [18a] Melnik, R.E., and Chow, R.: Asymptotic theory of two-dimensional trailing edge flows. Grumman Research Department Rep. RE-510 (1975). [18b] Measiter, A.F.: Boundary layer flow near the trailing edge of a fiat plate. SIAM J. Appl. Math. 18, 241—257 (1970). [19] Mekayn, D.; New inethods in lantinar boundary layer theory. London, 1961, [20] Niknradae, J.: Laminare Reibungarchichten an der lingaangestromten Platte. Monograph. Zontrale £. wiss. Berichtawesen, Berlin, 1942. [21] Prandtl, L.: Uber Fiissigkcitsbewegung bei schr kleiner Reibung. Proc, Third Intern. Math. Congr. Heidelberg 1904. Reprinted in: Vier Abhandlungen zur Hydro. und Aerodynamit, Gotlingon, 1927; NACA TM 452 (1928); ace also: Coll, Works 17, 675 —584 (1961). [22] Prandti, Ls ‘Phe mechanies of viscous fluids, In W.F. Durand: Aerodynamic Theory 117, 34—208 (1935). [23] Rotta, J.C.: Gronzachichttheorie zweiter Ordnimg fiir ebene und achsenaymmetriache Hyperschaliatromung. ZEW 15, 329—334 (1967). {24] Schmidt, H., and Schroder, K.: Laminare Grenzschichten. Bin kritischer Literaturbericht. Part I: Grundlagen dor Grenzachichttheorie. Luftfahrtforsclnng 19, 65—97 (1942). [25] Steinhener, J.: Die Lésungen der Biasinsschen Grenzachichtdifferentialgleichung. Proc. Wiss. Ges. Braunschweig XX, 96—125 (1968), [25a] Stewartson, K.: Multistructured boundary la; ‘ers on fiat plates and related bodies. Adv. Appl. Mech, 14, 146-239, Academic Press, New York, 1074, [26] Tollinion, W.: Gronzschichttheoric, Handbuch der Exper.-Physik 7V, Part 1, 241—287 (1931). [27] Topfer, C.: Bemerknagen 2 dem Anfitx vou TH. Blasina: Grenzachichton in Fliissigkeiten kl w Reibnng. Z, Math. Phys. 60, 307-~3098 (1912), [28] Woyl. H.: Concerning the differoutial equations of some boundary layer problems, Proc. Nab. Acad, Sci, Washington 27, 578 ~583 (1941). {29] Woyl, H.: On the differential equations of the simplest boundary layer problems. Ann. Math. 43, 381—407 (1942). {30} Janour, Z.: Resistance of a flat plate at low Reynolds anmbers, NACA TM 1316 (1951). CHAPTER VIII G eral properties of the boundary-layer equations Before passing to the ealculition of further examples of boundary-layer flow in the next. chapter, we propase first to discuss some general properties of the bound- ary-layer equations, In doing so we shall coufine our attention to steady, two- dimensional, and incompressible boundary layers, Although the botmdary-ltyer equations lave been simplified to a great extent, as compared with the Navier-Stokes equations, they are still so difficult from the mathematical point of view Ukat not very many general statements about them ean be made. ‘Ta begin with, it is important to notice that the Navier-Stokes equation of the elliptic type with respect to the coordinates, whereas Prandtl’s honndary-layer equations are parabolic. I is a cansequence of the simplifying assumptions in houndary-layer theory that the pressure can be assumed constant in a direction at right angles to the boundary layer, whereas along the wall the pressure can be regarded as being “impresserl” by the external flow so that it be- comes a given function, The resulting omission of the equation of motion perpen- dicular to the directiou of flow can be interpreted physically by stating that a fluid particle in the boundary layer has zero mass, and silfers no frictional drag, as far as its motion in the transverse direction is concerned. It. is, therefore, clear that with such fundamental changes introduced into the equations of motion we must, ate that their solutions will exhibit certain mathematical singularities, and that agreement: hetween observed and calculated phenomena cannot always be expected. a, Dependence of the characteristics of a boundary layer on the Reynolds numbert The assumptions which were made in the derivation of the boundary-layer equations are satisfied with an increasing degree of acenracy as the Reynolds number inereases. Thus boundary-layer theory cau be regarded as a process of asymptotic integration of the Navier-Stokes equations at very large Reynolds numbers*, ‘This statement leads us now to a discussion of the relationship between the Reynolds number and the characteristics of a boundary layer on our individual body under consideration, Lt will be recalled that in the derivation of the boundary-layer equations + Cf. Sees, VIEE and 1X}. * Tho argument contained in this soctiot was already discussed in See. VILE on higher-order approximations, The amplification ia given here for the anke of better underatnuding, a, Dependence of the characteristics of a boundary Inyer on the Reyuolds number 151 dimensionless quantities were used; all velocities were referred to the free-stream velocity Moo, all lengths having heen reduced with the aid of a characteristic length of the body, L. Denoting all dimonsionless magnitudes hy a prime, thus #[Us = Liye] = ; we obtain thie following equations for the steady, two-dimensional case: + On" au 8.1 w 4 (8.1) aul tie (8.2) y=0: wae yo=oo: WHU"(x'), see also eqs. (7.10) lo (7.12), Here R denotes the Reynolds number formed with the aid of the reference quantitios It is seen from cqus. (8.1) and (8.2) that the boundary-layer solution depends on one parameter, the Reynolds number R, if the shape of the body, and, hence, the potential motion U’(z’) are given. By the usc of a further transformation it is possible to eliminate the Reynolds number also from eqns. (8.1) and (8.2). If we put wav V~R= (8.3) and _ =y~Re= (8.4)t eqns. (8.1) and (8.2) transform into: wy ve (8.5) (8.6) with the boundary conditions: u’=0 and v” =0 at y” =0 and uw’ =U" at y" =o. These equations do not now contain the Reynolds number, so that the solutions of this system, i. e. the functions w’(2’, y) and v” (2’, y”), are also independent of the Reynolds number. A variation in the Reynolds number causes an affine trans- formation of the boundary layer during which tho ordinate and the velocity in the transverse dircction are multiplied by R-/2, In other words, for a given body the dimensionless velocity components 1/Us and (v/Ues) ‘(Uso L[v)¥? are functions of the dimensionless coordinates 2/f and (y/L) «(Veo L/¥)!/?; the functions, more- over, do not depend on the Reynolds number any longer. The practical importance of this principle of similarity with respect to Reynolds number consists in the fact that for a given body shape it suffices to find the solution to the boundary-layer problem only once in terms of the above dimensionless variables. transformation is identical with Uhat imptiod in eqns. (7.47) and (7.48). 162 VILL. Goneral propertios of the boundary-layer equations Such a solution is valid for any Reynolds number, provided that the boundary layer is laminar. In particular, it follows further that the position of the point of separation is independent of the Reynolds number. The angle which is formed between the streamline through the point of separation and the body, Fig. 7.2, simply de- creases in the ratio 1/R!/? as the Reynolds number increases, Morcover, the fact that, separation docs take place is preserved when the process: of passing to the limit R co is carried out, ‘Thus, in the case of body shapes which exhibit separation, the boundary-layer theory presents a totally different picture of the flow pattern than the frictionless potential theory, even in the limit of R00. This argument confirms the conclusion which was already emphatically stressed in Chap, TV, namely that the process of passing to the limit of frictionless flow must not be performed in the differential cquations themselves; it may only be under- taken in the integral solution, if physically meaningful results are to be obtained. b. ‘Similar’ solutions of the boundary-layer equations A second, and very important, question arising out of the solution of boundary- layer equations, is the investigation of the conditions under which two solutions are ‘similar’, We shall define here ‘similar’ solutions as those for which the com- ponent w of the velocity has the property that two velocity profiles u(, y) located at different coordinates x differ only by a scale factor in « and y. Therefore, in the case of such ‘similar’ solutions the velocity profiles u(x, y) at all values of x can be made congruent if they are plotted in coordinates which have been made dimen- sionless with reference to the scale factors. Such velocity profiles will also sometimes be called affine, The local potential velocity U(x) at section x is an obvious scale factor for u, because the dimensionless u(x) varies with y from zero to unity at all sections. The scale factor for y denoted by g(x), must be made proportional to the local boundary-layer thickness. ‘The requirement of ‘similarity’ is seen to reduce itself to the requirement that for two arbitrary sections, x, and x5, the components u(x, y) must, satisfy the following equation & {= [ylo(e)l} _ & {ear lyl9 (aa))} “T(%) Ul) (87) The boundary layer along a flat plate at zcro incidence considered in the preceding chapter possessed this property of ‘similarity’. The free-stream velocity Uso was the seale factor for w, and the scale factor for y was equal to the quantity g = V9 z/Uc which was proportional to the boundary-layer thickness. All velocity profiles became identical in a plot of u/UVa, against y/g =y¥ U.jyz = 1, Fig. 7.7. Similarly, the cases of two- and throe-dimensional stagnation flow, Chap. V, afforded examples of solutions which proved to be ‘similar’ in the present sense, ‘The quest for ‘similar’ solutions is particularly important with respect: to the mathematical character of the sohition. In eases when ‘similar’ solutions exist it ja possible, az wo shall see in more detail Inter, to reduoo the system of partial differentinl eqpintions to one involving ordinary differential equations, whieh, evi- dently, constitutes a considerable mathematical simplification of the problem. ‘The boundary laycr along a flat plate can serve as an example in this respect also. b, ‘Similar’ solutions of the boundary-layer equations 153 It will be recalled that with the similarity transformation n = y YU o/¥ x, eqn. (7.24), we obtained an ordinary differential equation, eqn, (7.28), for the streant function f(7), instead of the original partial differential equations. We shall now concern ourselves with the types of potential flows for which such ‘similar’ solutions exist, This problem was disenmcd in great detail first by S. Goldstein [4J, and. luicr by W. Mangler [9]. The point of d to consider the boundary-layer equations for plane steady flow, eqns, (7.10) and (7.11) together with eqn. (7.52), which can be written as ou ' atyn : | Ou ou au ou (8.8) wae thay RUT a | the boundary conditions being w == = 0 for y = 0, aud u =U for y= 00. ‘The equation of continuity is integrated by the introduction of the stream fauction pl, x) with ‘Thus the equation of motion becomes ay oy oy by day” dz (8.9) with the boundary conditions dy/ax = 0 and ayldy =0 for y = 0, and aylay for y == 00, In order to discuss the question of ‘similarity’, dimensionless quantit are introduced, as was done in Sec. VIIa. All longtis arc reduced with the aid of a suitable reference length, L, and all velocities are mace dimensionless with réference to a suitable velocity, Uso, As a result the Reynolds number p= lek u appears in the equation. Simultancously the y-coordinale is referred to the dimension- less scale factor g(x), 80 that we put. e f=f 2= a (8.10)+ } The transformation nt =a7" (L+24n), proposed by F. Schultz-Grunow [6a, [5a], makes it, possible lo reduce several problems in- volving self-similar solutions to that of the flat plate at zero incidence. If A = 8/2. is chosen as the curvature parameter, the transformations can he applied to flows along longitudinally curved walls with hlunt or sharp lending edges as well as with blowing ot auction (Chapt, XIV). ‘The preceding transformation is exact to second order in curvature which moans that all terms of the order A have been included. 154 VIII. General properties of the bonndary-layer equations ‘The factor VR for the ordiuate already appeared in eqn. (8.4). ‘The stream function is made dimensionless by the substitution HEN = “Luegiz),” (8.11) Conseqnently, the velocity components become us =i U x =Uf, (8.12) —VRenVh © Swat v9(%—n8 nf), where the prime in /’ denotes differentiation with respect to 7, and with respect, to x ing’. It is now seen directly from eqn profiles u(x, 7) are similar in the previously defined sense, when the atronm function f depends only on the one variable 4, eqn. (8.10), so that the dependence of f on € is cancelled, In this case, moreover, the partial differential equation for the stream function, equ. (8.9), must reduce itself to an ordinary differer equation for f(a). If we now procecd to investigate the conditions under which this reduction _of eqn. (8.9) takes place, we shall obtain the condition which must be satisfied by the potential flow U(x) for such ‘similar’ solutions to exist. If we introdiee now the dimensionless variables from eqns, (8.10) and (8.11) into equ, (8.9), we obtain the following differential equation for {(é, 4): ” " 0 U Lf i real BAA = 98 (PZ 2) (8.18) where and fare contynctions for the following functions of xt Jy é L , TUN: Bay’, (8.14) aud where (7 -:dU da, The boundary conditions far eqn. (8.13) are f ~ 0 and f= 0 for +: Oand f! == 1 far y jmilar’ solutions exist only when f and f' da not depend an & i.e, when the vight-liand side of eqn. (8.13) vanishes. Simullanconsly the coefficients « and f on the left-hand side of eqn. (8.13) must. be independent of x, ic. they must be constant, This latter condition, combined with eqn, (8.14), furnishes two equations for the potential velocity r) and tho scale factor g(x) for the ordinate, so that thoy can be evaluated. Hence, if similar solutions of boundary-layer flow are to oxist, the stream finetion {(77) must satisfy tle following ordinary differential ' equation: re vaff spa f%)-.0 (8.15) with the boundary conditions » 9. 20, fo; (8,16) b. ‘Similar’ sotution of the boundary-layer oquations 155 ‘This equation was first givon by V, M. Falkner and 8, W. Skan [2], and its solutions were lator studied in detail by D. R, Hartree [6], We shall revert to Uis point in the succeeding chapter. It remains now to determine from equ, (8,14) the conditions for U(x) and g(x). Prom (8.14) we obtain first Bafa iy GU) and hence, if 20 — B10, i ot =20—p 7. (8.17) Further from (8.14) we have L , a—p=y go U anu henee —p a gut OA yy PUG so that upon integration (a) =o" (8.18) where K is a constant. The elimination of g from equs. (8.17) and (8.18) yields the velocity distribution of the potential flow 2 = wea) iz It will be recalled that the ease 20 — f U au (8.19) cs and = 0 has been excluded. As scen from eqn. (8.14) the result. is independent of any common factor of a and B, as it can he included in g. Therefore as long as a +0 it is permissible to put « - -} 1 without. loss of generality. It is, furthermore, convenient to introduce a new constant m to replace B by putting mss : (8.21) so that, with « = 1, the velocity distrilution of the potential flow and the scale factor g for the ordinate become 156 VIL. General propertios of the boundary-layer equations U (itm (2 z\m 7% (itm i) , (8.22) 2 2Uy 9 Veit (8.23) and the transformation equation (8.10) for the ordinate is n= (8.24) It is thus concluded that similar solutions of the boundary-layer cquations aro obtained when the velocity distribution of the potential flow is proportional to a power of the length of are, measured along the wall from the stagnation point. Such potential flows ocour, in fact, in the neighbourhood of the stagnation point of a wedge whose inoluded angle is equal to x B, as shown in Fig. 8.1. It is casy to verify with the aid of potential theory that we have here U(e) =Ca, (8.25) where C is a constant. The relationship between the wedge angle factor B and the exponent. m is exactly that given in eqn. (8.21). a Gi Fig. 8.1. Flow past a wedgo. In the neighbour- a 76 hood of the leading edge the potential velocity distribution is U(z) = Cam oS Particular cascs for a == 1:(a) For B == | wehave m = |, and eqn. (8.22) becomes U(x) = ax. This is the case of two-dimensional stagnation flow, which was considered in See. Vb 9, and which led to an exact solution of tho Navier-Stokes equations. With & — 1, and B = 1, the difforential oquation (8.15) transforms into eqn. (5.39) which was already considered carlier. The transformation equation for the ordinate, eqn. (8.24), becomes identical with the already familiar equation (5.38), if we put Ufe =a. (b) For B == 0 we have m == 0, hence U(x) is constant and equal to U,.. theeaseof a flat plate at zero incidence. Wufollows from eqn. (8.24) that 9 =y V Ueof2 vx. ‘This value differs only by « factor V2 from that irttroduced in eqn. (7.24). Correspond- ingly the differential equation {’” +-{f" =0 which follows from eqn. (8.15) differs by a factor 2 in the second term from eqn. (7.28) which was solved earlier. The two equations become identical when transformed to identical definitions of 7. Solution for different. vatues of m will be considered later in Chap. 1X. a. Transformation of the boundary-layor equations into the heat-conduotion equation 157 The case a = 0: The case a == 0 which has, so far, been left out of account, leads, as is casily inferred from eqn. (8-19), to potential flows U(x) which are pro- portional to If for all values of 8. Depending on the sign of U this is the case of a two-dimensional sink or source, and can also be interpreted as flow in a divergent or convergent channel with flat walls. This type of flow will also be considered in greater detail in Chap. LX. The second case excluded earlier, namely that when 2a — B == 0, leads to ‘similar’ solutions with U(x) proportional to e, where p is a positive or negative constant. We shall, however, refrain from discussing this case. . ‘The problem of the existence of similar solutions involving non-steady boundary layers was discussed by II. Schuh [15]; the same problem in relation to compressible boundary layers will be discussed in See, XIIId. 4. Transformation of the boundary-layer equations into the heat-conduction equation R. von Mises [10] published in 1927 a remarkable transformation of the boundary-layor oquations. This transformation exhibits the mathematical character of the equations even more clearly than the original form. Instead of the Cartesian coordinates x and y, von Mises introduced the stream function y, together with the length coordinate x as independent variables. Substituting into eqns. (7.10) and (7.11), as well as introducing the new coordinates & = x and 1 = p instead of x and y, we obtain au du 2E | dudy du du ae = 26 a2 t Bq de = 5E Yay’ au 9426 4 20m gy 2 ay = 8 By + aq dy —9 + ay” Hono, from oqn. (7.10), it follows that au, .dp_ a (aw wot eed (ul). Introducing, further, the ‘total head” g=pt4ou, (8.20) where the small quantity 4 gv? can be neglected, we obtain, reverting to the symbol q Y20 8 zx for €: ag ar (8.27) Viren. We may also put 158 VIIT. General proporties of the boundary-layer equations Equation (8,27) is a differcutial equation for the total pressure g(x, y), and its honndary conditions are g =pla) for y=0 and g == p(x)-+ } QU? = const for p =00. In order to represent the flow in the pl from yp to y with the aid of the equation = f 7 = Vif FooG quation (8.27) is related to the heat-conduction equation, ‘The differential equation for the one-dimensional case, ¢. g. for a bar, is given by Ls al a oxt’ ral plane x, y, it is necossary to transform (8.28) where 7' denotes the temperature, ¢ denotes the time, and a is the thermal diffusivity, sce Chap. XII, However, the transformed boundary-layer equation, unlike eqn. (8.28), is non-linear, because the thermal diffusivity is replaced by yu, which dopends on the independent variable x, as well as on the depend At the wall, y =0, « = 0,9 = p, eqn. (8.27) exhibits an unpleasant singularity. The left-hand side becomes ag/ax == dpjdx +0. Ou the right-hand side we have 1 = 0, and, therefore, 2g/0y = oo. 'This circumstance is disturbing when numerical methods are used, and is intimately connected with the singular behaviour of the volocity profile near the wall. A detailed discussion of eqn. (8.27) was given by L. Prandtl [LE], who had deduced the transformation a long time before the paper by I. von Mises appeared, without, however, publishing itf, ef. [1, 12, 16]. variable g. NJ, Luckert [8] applied oqn. (8.27) to the example of the boundary layer ona flat plate in order to test its practicability. L. Rosenhead and H. Simpson [13] gave a critical discussion of the preceding publication. c. The momentum and energy-integral equations for the boundary layer A complete calculation of the boundary layer for a given body with the aid of the differential equations is, in many cases, a3 will be scen in more detail in the next chapter, so cumbersome and time-consuming that it ean only be carried out with the aid of an electronic computer (sce also See. 1X i), Te is, thevefore, desirable to possess at least approximate methods of solution, to be applied in cases when an exact solution of the boundary-layer equations cannot be obtained with a reasonable amount. of work, even if their accuracy is only limited. Such approximate methods can be devisecl if we do not. insist on satisfying the differential equations for every fluid particle. Instead, the bonncary.layer equation is satisfied in a strataum near the wall and near the region of transition to the external flow by satisfying the boundary f See footnote on p. 79 of rof, [11] and the letter of L. Pranrltl to ZAMM 8, 249 (1928), ¢. The momentin and energy-integral equations for the boundary layer 159 conditions, together with certain compatibility conditions. In the remaining region of fluid in the boundary layer only a mean over the differential equation is satisfied, the mean being taken over the whole thickness of the houndary layer. Sich a mean yalue is obtained from the inomentiun cquation whiel: is, in turn, derived from the cquation of motion by integration over the bonndary-hyer thickness. Sinco this equation will be often used in the approximate methods, to bo disenssed later, we shall deduec it now, writing it down in its madern form. The equation is known as the momentum-integral equation of boundary-layer theory, or as von Kiirméin’s integral equation {7}. We shall restrict ourselves to the case of ateady, tva-dinensional, and incom- pressible flow, i.e. we shall refer to eqns. (7.10) to (7.12). Upon integrating the equation of motion (7.10) ‘vith respect to y, fram y == 0 (wall) to y =k, where the layer y =: h is everywhere ontside the boundary layer, we obtain: s ou _% 20 / (« as ox —U say = . (8.29) yoo ‘The shearing stress at the wall, zo. has been substituted for j(2u/éy)g, so that eqn. (8.29) is seen to be valid both for laminar and turbulent flows, on condition that in the latter case u and » denote the time averages of the respective velocity y components. The normal velocity com pouent, », can be replaced by v = -- f (Aujax)dy, 0 as seen from the equation of continuity, and, conseqnently, we have » [bay 08) yet Hitegrating by parts, we obtain for the second term 4 ‘ au [@/% ca)anof say [x5bov, ae co that. A au au To J (i - 0x8 2, é which ean be contracted to ; a AU ; “ a a a = (8.208 fae u)|dy + as ay dy 2, (8.29a) 3 d Since in both integrals the integrand vanishes outside the boundary layer, it is permissible to put k 00 , We now introcluec the displacement thickness, 6,, and the momentum thickness, 5,, which have already been used in Chap. VII. They are defined by 160 VIII, General properties of the boundary-layer oquations 6, U = f (Uw dy (displacoment thickness) , (8.30) and ad 5b, U2 = fuo—w dy (momentum thickness) . (8.31) yao It will be noted that in the first term of the eqn, (8.29a), differentiation with respect to x, and integration with respect to y, may be interchanged as the upper limit A is independent of 2. Hence d dU = gz (U7 4) +4, U Ge |. (8.32) This is the momentum-integral equation for two-dimensional, incompressible boundary layers, As long as no statement is made concerning tp, eqn. (8.32) applies to laminar and turbulent boundary layers alike, This form of the momentum integral equation was first given by H. Gruschwitz [5]. It finds its application in the approximate theories for laminar and turbulent boundary layers (Chaps. X, XI and XXII). Using a similar approach, K, Wicghardt [17] deduced an energy-integral equation for laminar boundary layers, ‘This equation is obtained by multiplying the equation of motion by u and then integrating from y = 0 to y= h > 4(z), Substituting, again, v from the equation of continuity we obtain . , bu Bu Ou au ef [wa —¥5 (J Bey) ~ +0 te] y= ! d ‘The second term can be transformed by integrating by parts: a bu 1 eu [hal {eelvn af oes, whereas by combining the first with the third term we have A au av 2% yt / [w a "YU ae é Finally, upon integrating the right-hand side by parts, we obtain h [Flt ray. a $ e ig fore dy uf Fy ay. (8.33) é é ‘Tho uppor limit of integration could here, too, be replaced by y = 00, because the integrands become equal to zero outside the boundary layer. The quantity p(du/ay)? represents the energy, per unit volume and time, which is transformed into heat hy friction (dissipation, of. Chap. X11). The term § 9 (U2—u2) on the left-hand d. The momentum and energy-integral equations for tho boundary layer 161 side represents the loss in mechanical energy (kinetic and pressure energy) taking place in the boundary layer as compared with the potential flow. Hence the term he f u(U—u%) dy ropresonts the flux of dissipated energy, and tho left-hand ide ° represents the rate of change of the flux of dissipated energy per unit, length in the x-direction. Tf, in addition to the displacement and momentum thickness from eqns. (8.30) and (8.31) ‘espectively, we introduce the dlissipation-energy thickness, ,, from the definition U8 by = [ u(U?—u%) dy (energy thickness), (8.34) é we can rewrite the energy-integral equation (8.33) in the following simplified form: rp 2 it (U4) = ay f (2) dy (8.35) i which represents the energy-integral equation for two-dimensional, laminar boundary layers in incompressible flow +. In order to visualize the displacement thickness, the momentum thickness, and the energy-dissipation thickness, itis convenient to ealenlate them for the simple case of linear velocity distribution, as shown in Fig. 8.2. In this case we find: displacement thickness 3, = $6 momentum thickness bn ~=h6 energy thickness dy == 0. ‘The extension of the preceding approximate method to axially symmetrical boundary layers will be discussed in Chap. XL. Approximate methods for thormal boundary layers are treated in Sec. XIIg; those for compressible and non-steady boundary layers will be given in See, XIIfd and Chap. XV, respectively, Fig, 8.2. Boundary layer with linear velo- city distribution 4 bounaaey-layer thickness 4, = displacement thickness 2, — evomentaim thtekness 4, — eneray thickness + In the case of turbulent flows, the energy-integral equation assumes the form a os tu ae (PH)=2 f THY 162 VIII. Goneral properties of the boundary-layer equations References, (1) Betz, A.: Zur Berechnung des Uberganges laininarer Grenzschichton jn die Aussenstromu Fifty years of boundary-layer research (W. ‘Tollion and H. Gorter, ed.), Brannsch wei 1955, 63—70. {2] Falknor, V.M., and Skan, S,W.: Some approximate solutions of the boundary tayer equa. tions, Phil. Mag. 12, 865—896 (1931); ARC RM. 1314 (1930), [3] Gois, ‘Th.: Ahntiche Grenzschichten an Rotationskirporn. Fifty years of boundary layer research (W. ‘Tollmien and HH. Gértler, ed.), Braunschweig, 1955, 294—303. [4] Goldstein, S.: A note on the boundary layer equations. Proc. Cambr. Phil. (1939), [5] Gruschwitz, F.: Die turbulonte Reibungsschicht in ebener Stré Druckanatieg. Ing.-Arch. 2, 321—346 (1931), [6] Hartreo, D.3.: On an equation occurring in Fatkher and Skan's approximate treatment of the equations of the boundary layer. Proc. Cambr. Phil. Soc. 33, Part 11, 223 — 239 (1937). {6a} Holt, Mu: Besie developmenta in itiid dynamies. Contribntion of It. Sehultz-Grunow and W. Breuer, 377—436, New York, 1965. [7] von Kérinén, Th.: Ober laininare und turbulente Roibung. ZAMM /, 233--253 (1921). Engl. transl. in NACA TM 1092; soe alo Coll. Works 17, 70-97, London 1956. [8] Luckort, H. J.» Uber die Integration der Differentialgloichung oiner Gleitschicht in ziher Flissigkeit, Diss. Berlin 1933, roprinted in: Sehriften dea Math. Sominars, Inst. £ angow. Math. dor Univ. Borlin 1, 245 (1933). [9] Mangior, W.: Die “ahnlichen” Losungeh der Prandilschen Grenzachichtgicichn 23, 241-261 (1943). [10] von Mises, R.: Bemerkungen zur Hydrodynamik. ZAMM 7, 425-431 (1927). (11) Prana, 1.: Zur Berechnung der Grenzachichten. ZAMM 13, 77—82 (1938); see also Coll, Works 17, 663-672, J. Roy. Aero. Soc. 45, 35~40 (1941), and NACA TM 959 (1940). [12] Riogols, I”, and Zaat, J.: Zam Obergang von Grenzschiehten in die ungestarte Strénung, Nachr, Akatl. Wiss. Gottingen, Math. Phys. Klasse, 42-45 (1947). [13] Rosenhead, L., and Simpson, J.H.: Note on the velocity distribution in the wake behind 8 fiat, plate placed along the stream. Proc. Cainbr. Phil. Soe. 32, 285-291 (1936). [14] Schréder, K's Verwendinig der Diflerenzenrechnung zur Boroohnang der laininaren Gronz- schicht, Math. Nachr. 4, 439—467 (1951). (15] Schuh, H.: Ober dic “ahnlichen’” Lasungen dor instationéren Inminaren Grenzschicht- gleichang in inkompressibler Strémung. Fifty yoars of boundary-layer research (W. Tollwien and H. Gartler, ed.), Braunschweig, 1955, 147—152. [15] Schultz-Grunow, IF., and Henscler, TH. che Grenzachichtloaungen zweiter Ordaung fir Stromungs- 1nd “Vemperaturgrenzschichten an longitudinal gekriinmten Wiinden mit Gronzachichtbecintlussung. Wérme. und Stoffithertragung 7, 214—219 (1968). [16] Tollmien, W.: Uber dag Verhalten ciner Stromung lings ciner Wand am fusseron Rand ihror Reibungsschicht, Botz, Anniversary Volume, 218 ~224 (1945). [17] Wieghardt, I: Ubor cinen Knergicsatz, zur Borechnung laminarer Grenzachichten. Ing.. Arch, 16, 231 - 242 (1948). Soe. 35, 338—340 jung bei Druckabfall und CHAPTER IX Exact solutions of the steady-state boundary- layer equations in two-dimensional motion ‘The present chapter will deal with some exact solutions of the boundary-layer equations. A solution will be considered exact when it is a complete solution of the boundary-layer equations, irrespective of whether it ia obtained analytically or by numerical methods. Ou the other hand, Chap. X will deal with approximate solutions, i.e, with solutions which are obtained from integral relations, such ns the momentum and energy-integral equations «eseribed in the preceding chapter, rather than from differential equations, ‘There are in existence only comparatively few exact analytical solutions, and we shall discuss them first. Generally speaking, te process of obtaining analytical solutions of the boundary-layer cquations encounters considerable math difficulties, as alreasly illustrated with the example of a flat plate. equations ure non-linear in most. eases so that, again generally speaking, they be solved only by power-series expansions or by numerical incthods. Even for the physically simplest case of the boundary layer on a flat plate at zero incidence with incompressible flow no closed-form analytic solution haa been discovered so far. mensional motion, the bounda In the case of two-« ‘ (7.10) to (7.12): their boundary conditions are given by eq ver equations amd (0.1) au, au a tu me Hy Wy 9.2 woe bay OU ae ay?? (02) y=0; w=0, v=0; yoo: w= U(x). (9.3) 0. a wa, pn az Consequently the atream function must. satisfy the following equation [sco also equ. (7.18)]: ; oy oy op oy 7, dU ay 9.4 dy dzay ~~ ox byt =U de +” ays? (9.4) 164 IX. Fxect solutions of the steady-state boundary-layer equations. with the boundary conditions ay/ay = 0 and dy/dx = 0 wt the wall, y = 0, and dyldy = U(x) at y = 00, a. Flow past a wedge The ‘similar’ solutions discussed in Chap. VIL constitute a particularly simple class of solutions «(a,y) which have the property that the velocity profiles at different distances, z, can be made congruent with suitable scale factors for w and y. The systers of partial differential equations (9.1) and (9.2) is now reduced to one ordinary differential equation. It was proved in Chap. VIET that such similar solutions exist when the velocity of the potential flow is proportional to a power of the length coordinate, z, measured from the stagnation point, i. e. for U(2) =u a". From eqn, (8.24) it follows that the transformation of the independent. variable y, which leads to an ordinary differential eqnation, is: : +1U a a ae 2 ‘The equation of continuity is integrated by the introduction of a stream function, for which we put + (9.5) n met van=Vaaq vem 2° i, ™ as acon from eqns. (8.11) and (8.23). ‘Thus the velocity components become um, 2 f'(n) =U f'(n), PEP owe estar}. Introducing these values into the equation of motion (9.1), dividing by m 1, 22"- and putting, as in eqn. (8.21), (9.6) : (9.7) m= we obtain the following differential equation for f(7): + ff" + Bf?) =0. (9.8) Ut will be regaled! that it was already given as ci, (8.15), and that its boundary conditions are 70: [=0; moo: fal. Kqnation (9.8) was first deduced by V.M. Falkner and §. W. Skan, and its solutions were later investigated in detail by D. R. Hartree (see References to Chap. VIIL). ‘The solution is represented in Fig. 9.1. In the case of accelerated flow (m>0.f >0) a. Flow past a wedge 165 the velocity profiles have no point of inflexion, whereas in the case of decoleratod flow (m <0, 2 <0) they exhibit a point of inflexion. Separation occurs for 6 = — 0-199, ic. for m = — 0-091. This result shows that the laminar boundary layer is able to support only a very small deceleration without separation occuring. K. Stewartson [64] gave a detailed analysis of the manifold of solutions of eqn, (9.8). According to this analysis, in the rango of increasing pressures (0.199 + < B <0) there exists a further solution, that is, in addition to the one discovered by Hartree, The additional solution leads to a velocity profile with back-flow (cf. Chap. Xf). The potential flow given by U(x) = 1, 2™ existe in the neighbourhood of the stagnation point on a wedge, Fig. 8.1, whose included angle B, is given by eqn. (9.7). Two-dimensional stagnation flow, as well as the boundary layer on a flat plate at zero incidence, constitute particular eases of the present solutions, the former for B =1 and m =I, the latter for B =0 and m =0. 10 « uh) 08 Fig. 9.1. Velocity distri- 06 bution in the laminar . boundary layer in the flow pete wetlge given 04 by U(x) = u,2™. The exponent mand the wedge angle B (Fig. 8.1) O2 are connected through eqn. (9.7) ‘The caso f == 4, m is worthy of attention. In this case tho differential equation for f(y) becomes: f’” -- ff" | 4(L—f'2) = 0; it transforms into the differential equation of rotationally syminctrical flow with stagnation point, oq. (5.47), ie, g'" +2969" +1 —$' =0 for J(2), if we put » = V2 and dffdy = dd/ac. ‘This means that the calculation of the boundary layer in the rotationally symmetrical case can be reduced to the calculation of two-dimensional flow past a wedge whose included angle is xB = x/2. The relationship between the two-dimensional and rotationally bymmetrical boundary layers will be further 1) with amaximum near the wall are admitted. In sucht cakes, the Himit (9) = 1 for 9 > co i8 attained asymptotically “from above" rather than “from below" as was the care a0 far, Stclt solutions can he interpreted physioally as correspondalg to a laminar wall-jet produced in an external stream with a positive pressure gradient, dp/da > 0. Reference [63] demonstrates that the limiting caro of thero solutions, obtained when the maximum velocity excess tents to infinity, transforms into the well-known self-similar solution of a pure wall-jet in the absence of an external velonity — a cnsc tronted hy M, B. Ghutert (ace [40] in Chap, XI) ~ when we pnt p = —2. A particularly detailed monograph on exact, self-similar solutions for laminar boundary layers in two-dimensional and rotationally symmetric arrangements, inchtsive of the associated thermal boundary layers (see Chap. XII).was published by C.F. Dewoy and J. F, Gross (14). ‘Their considerations include the effects of compressibility (nee Chap. X11) with and without heat transfer, relate to varying vahtes of the Prandtl number, and inelude some cases of suction and blowing, K. K. Chen and P. A. Libby [9] carried out an extensive investigation of boundary Inyers which are characterized by smail departures from the self-similar wedlgo-flow houndary layers of the Falknor-Skan type. Evidently, sch boundary layers are no longer self-similar. b. Flow in a convergent channel ‘The case of potential flow given by the equation U(x) = — (9.9) = related to flows past a wedge, and also leads to ‘similar’ solutions, With u, > 0 it represents two-dimensional motion in a convergent charnel with flat walls (sink). The volume of flow for a full opening angle 2x and for a stratum of unit height is Q@ = 22 u, (Vig, 9.2). Introducing the similarity transformation ec 5 (9.10) as well as the stream funetion (zy) Fig. 9.2, Flow in a convergent channel b. Flow in a convergent chanel 167 we obtain the velocity components waUf;v=—Yrm ty. (9.11) Substituting into equ, (9.2) we obtain the differential equation for the stream finetion fr —ft+i=o. (9.12) The homulary conditions follow from eqn. (9.3) aud are: f> Oat O,mad fF and f= Oak n == 00, ‘This is also a particular caso of the class of ‘similar’ solutions considered in Chap, VILL, Equation (9.12) is obtained from the more geucral difleren- tial equation (8.15) for the ease of ‘similar’ boundary layers, if we put a = 0, and f= + 1. ‘The example under consideration is one of the rare cases when the solution of the bonndary-lnyer equation can be obtained analytically in closed form, First, upon multiplying eqn. (9.12) by /” and integrating once, we have fe —R(f lf +2) =a, where @ is a constant of integration, Ils value is zero, asf’ = 1 and f” ~0 for 7 +00. Thus 2 er +2) a i af V3 [estas where the acditive constant of integration is seen to be equal to zero in view of the honndary condition f’ = | at 7 = 00, ‘The integral can be expressed in closed form as follows: _ n=V2 { tant VEL — tanh? V3} , or or, solving for f’ = uf: f= v = 3 tanh? (i + 140) —2. (9.13) There we have substituted tanh-t y/3 as well as Q = 227 U(r = radial dist ogn. (9.10) by 1.146, Tutroducing the polar angle 0 > yf, nce from the sink), we ean replace 1 from (9.14) ‘The velocity distribution given by eqn. (9.13) is represented in Fig, 9.3. Abn = 3. approximately, the boundary layer merges with the potential flow, Monce the boundary-layer thickness becomes} == 32 )/9/T/r ; it decreases, asin other examples, as IVR. 168 1X, Exact. solutions of the steady-state boundary-layer equations Fig. 9.3, Velocity distribution in the Jaminar boundary Jayer of the flow in a convorgent channel 40 ove ‘The preceding solution was first obtained by K. Pohlhawsen [50]. It will be recalled from See. Vb E2 on p. 107 that the flow through a divergent channel dis- cussed by G. Hamel constitutes an exact solution of the Navier-Stokes equations. ‘The diagram in Fig, 5.15 contained some mimcrical results pertaining to this solution. In this connexion, a paper by B. L, Reeves and C. 4. Kippenhan [52] may usefully be consulted. c. Flow past a cylinder; symmetrical case (Blasius series) ‘The class of ‘similar’ solutions of the boundary-layer equations considered so far is comparatively narrow. Apart from the examples of the flat plate, stagnation flow, flow past a wedge, anc flow in a convergent channel which have already been described, few aclditional solutions ean be obtained. We shall now consider the general case of the boundary layer on a eylindrical body placed in a stream which is porpen- dicular to its axis, The method of solution was first given by H. Blasius [4]; it was developed further by K. Hiemenz [39] and L. Howarth [40]. It is necessary to distingnish here two cases depending on whether the eylinder is symmetrical about an axis which is parallel to the stream at a large distance from the body or not. We shall vefer to these two cases as to the symmetrical and asymmetrical case, respectively. In either case the velocity of the potential flow is assumed to have th of a power scrics in a, where x denotes the distance from the stagnation point mensured along the contour, ‘The velocity profile in the boundary layer is also represented as a similar power series in x, where the coefficients are assumed to be functions of the coordinate y, measured at right angles to the wall (Blasius series). L, Howarth succeeded in finding a substitution for the velocity profile which confers universal validity on the y-dependent coefficients. In other words, by a suitable anaunption regarding the power series, ils coefficients have been made independent of the particulars of the cylindrical body, so that the resulting functions could be evaluated and presented in the form of tables. Thus the ealeulation of the boundary imple if use is nade of the tables, provided iently large number of torms of the series, layer for a given shape becomes very that the tabulation extends over a suff ‘The usefulness of Blasina’s incthod is, however, severely restricted hy the fact that, precisely in the inost important case of very slender body-shapes, a large ¢. Flow past a cylinder; symmetrical case ( Blasius acries) 169 number of terms is requ ; in fact, their number is so large that it ceases to be practicable to tabulate them all with a reasonable amount of numerical work, This ia caged! by the cirenmnstance that in the case of slender boty-sections, ¢. g. in the case of an cllipse, placed in a stream parallel to its major axis, or in the ease of ai aerofoil, the potential velocity near the stagnation point in the neighbourhood of the Ieacling edge increases steeply at first anc then varies very slowly over a consi- ble distance downstream. A function of this type cannol be well represented by 2 power series with a small number of terms, In epite of this limitation Blasiua’s method is ol great fundamental importanee because, in cases when its convergence is insufficient to reach the point of separation, it can be used to calculate analytically ancl with great acenracy the initial portion of the boundary layer near the stagnation point. The calculation can then be continued with the aid of a suitable numeri integration method, such as, for example, the one described in Sec. LXi. We shall now very briefly describe the procedure that is followed for the caleu- lation of a boundary layer with the aid of a Blasius scries. A more detailed account can be found in the earlier editions of this book [57a]. However, the numerical results for the circular eylinder are given more fully. We consider the symmetric case anc assuine that: the potential flow the form of the series givon in U(x) = rt 4 tg x3 tg ad . (9.15) The coefficients w, us, ... depend only on the shape of the body and are to be con- sidered Inown. The continuity equation is satisfied by the introduction of a stream- function y (x, y)- In analogy with eqn. (9.15) it is plansible also to adopt a power series in #, its coefficients being treated as functions of y. ‘The choice of the particular form of the power series is governed by the desire to render the funetions of y con- tained in it independent of the coefficients my, ug, us, . . . which describe the potential flow, In this manner, the functions of y become universal and can be calculated once and for all. ‘The distance from the wall is made dimensionless by assumingt y y . (9.16) } (9.17) n This leads to the form = Ve (04 2 fil) +4 ts 2° fol) + 6 u5 28 fal) + «- for the stream-function with the aid of which it now becomes possible to determine the appropriate series for the velocity components u = dp/dy and v = — dy/dz, Substituting these expressions into the equation of motion (9.2) we compare coefli- ig in it the first, term (9.15). i. 6, aa for Un. This brings with it the disadvantage that it docs not make an allowance for the increase in the houndary-layer thickness in the dowustream dircet 170 IX. Exact solutions of the steady-state boundary-layer equations cients and thus obtain a systcin of ordinary differential equations for the functions hy fa... .. The first two equations turn out to be 2 Aha 14h’ th h— 3h fe his = 1 bs Tn these, differentiation with respect to 7 is denated by primes, The agsooitted boundary conditions are n= Oh neo: finl f=. (9.18) (9.19) All differential equations for the functional coefficients are of the third order, and only the first one, that. for fy, is non-linear; it. is identical with the equation for two- dimensional stagnation flow, eqn. (6.39), discussed in Chap, V, All remaining equa: tions are linear and their coefficients are expressed in terms of the functions asso- ciated with the preceding terms. The functions f; and fz have been calculated already by K. Iiomenz [39], and their first derivatives are represented graphically in Fig. 9.4, The function f; for the velocity distribution was reproduced earlier in Fig. 5.10 and Table 5.1 (when it wag denoted by $/). The higher-order functions can he found in the earlier editions [57 a}. 15 10 08 Vig. 9.4, The functions /,’ and /,” which appear in the Blasias power seri base the calculation on the pressure distribution obtained from potential theory, although in the literature the problem was frequently solved with the aid of an experimentally determined pressure distribution. ‘Phe ideal yelocity distribution in scours, irrotational flow pasta circular cylinder of radius Rand free-stream velocity Um ynuallel to the x-axis is given hy | u(x) = 2 Un sin afk 2 Uesing, (9.20) whore @ is the angle measured from the stagnation point. Expanding sin a/R inta a series and comparing it: with Uhut in eqat. (9.15), we find that Un ¥ Bre UnR and g = e. Flow past a eylinder; syminetrinal case (the Blasius series) 171 3 ae Fig. 9.5. Velocity distribution in the boundary layer on a cireular cylinder 6 —angle measured from stagnation point 0 ao ae a4 06 08 0 he to The velocity profiles for different. values of # are seen plotted in Fig. 9.5 which is based on a sories for the velocity, 2, carried as far as the term in 2!!. The velocity profiles for # > 90° possess a point of inflexion because they lic in the region of increasing pressure. The distribution of shearing stress t) = p1(0u/2y)o is plotted in Fig. 9.6. The position of the point of separation results from the condition that to = 0, and is given by $5 = 108-8° 4 Fig. 9.6. Variation of shearing stress nt 70U8/ * the wall over the circumference of a cit- der for a laninar boundary 0 cular layor 172 1X. Fxact solutions of the steady-state boundary: er equations If the power sorics were terminated at x9, the point of separation would turn out to he at $s = 109-6°. Better acenracy can nowadays be obtained with numerical methods, see Seca. 1Xi and Xo3. The accuracy of this calculation based on a power acrica can be tested for speed of convergence of the omitted portion of the series by invoking the conditions of compalibility at the wall. According to eqn. (7.15), we must have pat dz (9.21) Verification of the first com- tion from eqn. (9.21) for the laminar boundary layer on a circular cylinder from Fig. 9.5. The first, compatibility condition is sat fied approximately as far ax some point beyond separation Figure 9.7 compares the curvature of the velocity profiles measured at the wall with its oxact value represented by UdUj/dx. The agreement is good for a distance x beyond the point of separation, We may, therefore, conclude that the Blasins series ating at the tern at! satisfies the compatibility condition on a circular cylin- der up to a point which lies beyond the point of separation. It does not, however, necessarily follow that. the Luneated series represents the velocity profile with good acenrney. As already mentioned, in the case of more slender body-shapes considerably more terms of the Blasins serics are required, if it is desired to obtain the velocity profiles as far as the point of separation. However, the cvahiation of further functional coefficients is hindered by considerable difficulties. These are due not only to the fact. that for every additional term in the scries the number of differential equations to be solved increases, but also, and even more forcibly, the difficulties are duc to the need to evaluate the funétions for the lower power terms with ever increasing aceuracy, if the functions for the higher power terms are to be sufficiently accurate L. Howarth [40] extended the present method to include the asymmetrical case, but the tabulation of the functional coefficients was not carried beyond those corresponding to the power ?. N. Froessling [23] carried out. an extension of this method to the rotationally symmetrical ease which will be considered in Chap. XL Measurements of the pressure distribution around a circular cylinder were reported hy KK. Hicmenz in his thesis presented to Gocttingen University [39]. ‘They d. Boundary layer for the potential flow given by U(z) = Us — ax" 173 were made the basis of his boundary-layer caleulotions. Ilis measurements showéd separation at ds = 81°, whereas the calculation indicatedif, = 82°. Later 0. Flachs- bart. published extensive experimental data an the preasure distribution, Fig. 1.10, which point to a large influence of the Reynolds number. For values of Ue Reynolds nuinber below the critical the pressure minimum occurs nlready near 6 ~~ 70°, wnd the pressure is nearly constant over the whole dowuslreant portion of the cylinder, Vor Reynolds numbers above the critical the pressore minimnin shifts to 6 = 90° approximatsly, in agreement with the potential-flow theory and, on the whole, the pressure distribution departs less from that given by the potential theory than in the previous case. Between these values, i. c. near a critical Reynolds number of approxi- mately Uso D/v = 3 x 105, the drag cocflicient. of the circular cylinder decreases abruptly (Fig. 1.4), and this phenomenon indicates that the boundary layer has beeome turbulent (see See. XVEIIf). The laninar boundary layer on a circular cylinder was also investigated by A. Thom [67], at. a Reynolds number Uz. D/v = 28,000 and by A. Fage [16] in the range Us Div = 10 to 3-3 x 10% A paper by L. Schiller and W. Linke [54[ contains some considerations concerning pressure drag and skin friction in the region of Reynolds numbers below the critical. In the range of Reynolds numbers: fron: about 60 to about 5000 there exists behind the cylinder a vortex street which shows a regular, periodic structure (Figs. 2.7 and 2.8). The frequency at. which vor- tices are shed in this so-called von Karman vortex strect has been investigated by IL. Blenk, D. Fuchs and H. Liebers, and, more recently by A. Roshlo (see Chap. 11). 4. Boundary layer for the potential flow given by U(x) = Uy ~ az” A further family of solutions of the boundary-layer equations was found by L. Howarth [41] and I. Tani [66]. These solutions relate to the potential How given by U(z) = Uy —azn (n= 1,2,3...), (9.22) which, evidently, constitutes a generalized forin of the flow along a fiat plate (sce Sec. Vile), aiid becomes identical with it when we put a = 0. In the simplest case witlt n = 1, which was treated by L. Howarth, tho flow ean bo interpreted as that which ocours in a channel which consists of a portion with parallel walls (velocity Us) followed by cither a convergent (a < 0) or adivergent (a > 0) section. This is another example of a boundary layer for which the velocity profiles are not similar. L. Howarth introduced the new independent, variable 1 4/Uo . <9 9.23 ” av Vos (9.23) which is identical with that used in the flat plate solution at zero incidence. He assumed further (x* <0, accelerated flow; z* > 0, decelerated flow). Itis now possible to stipulate a power series + When equation (9.22) is written in the form U(r) = U,(1—z/L) for n = 1, it can also be interpreted as representing the potential low along a flat wall which starts at « = 0 and which ahuts on to anothor iufinite wall at right anglos to it at 2 = Z. It is of the same type as Ue cane of decelerated stagnation flow shown in Fig. 2.17, the stagnation point being at x = D. 174 TX. Exact solutions of the steady-state boundary-layer equations in 2* for the stream function in a manner si fioionts being funetions of y: ilar to the case of the cylinder, See. 1X, the coef. lay) = VUs ve (lon) — (8 2*) f(r) + (8 z*) faln) — + - (9.24) Hence the velocity of flow becoines = Bo f{lo'(n) — (8 2") fy'(m) + (8 2") f'(n) — + - (9.25) Introducing these values into the equations of motion (9.2) aud comparing coefficionta we ohtain a syatein of ordiunry differential equations for the functious {o(2), /,(7), -..- The first three of these are: Io" + Iolo” = 0, WO * fo” 2h ht fo" = —2, I+ Iola” —4 fol Id +5 fo le b+ 2TH, with the houndary eonditiona n=O: fo =hi=0: nao fl ms Only the first equation is non-linear, and it is identical with that for a fiat plate at zoro inci- donce 3. All remaining equations are linear and contain only the function /, in the homogeneous portion, whoreas the non-homogeneous terms are forincd with the aid of the remaining functions /,- L. Howarth solved the first seven differential equations (up to and ineluding /,), and cal- culated tables for them, ‘The series (9,25) converges well with these values of /, in the range — 0-1 0) the point of separation ix at x* = 0-12 approximately, but for the slightly extended range of values the convergence of the series (9.25) is no longer assured. In order to reach the point of separation, L, Howarth used a namerical procedure for the continnation of the solntion, Velocity profiles for several valnes of x* for both accelerated distribution in the cr Howarth [41] Fig. 9.8. Veloci ar houndary Inyer for the potcutial flow given by U() = Uy are + The independent variable 7 in the above cquations differs from that in Chap. VIT by a factor }, ¢. Flow in the wake of flat plate at zero incidence 175 and deeclorated flow are acest plotted in Fig. 9.8, It. should be noted that: all profiles in decelerated flow have a point of inflexion. D. It. Hartroe [38] repented these caloulations and obtained good agreement with 1. Howarth. The case for a/Uo = 0-125 was calculated more acon hy D. Leigh (44] who used an electronic digital computer for the purpose and who paid special attention to the region of separation. ‘The value of the form factor at the point of reparation itself was found to be 2* = 0-1198. ‘The method employed by L. Howarth was extended by 1. 'Taui [60] to include the cases corresponding tom > I (with a > 0), However, l. Tani did not publish any tables of the fi tional coefficients but confined himself to reporting the final result for n = 2, 4 and 8, fi his case, too, the poor convergence of the series did not perinit him to determine the point of separ: tion’ with sufficient acenracy and he found himself compelled to use 1. Howarth's nmnerieal continnation seheme, e. Flow in the wake of flat plate at zcro incidence ‘The application of the boundary-layer cquations is not restricted to regions near a solid wall. They can also be applied when a stratum in which the influence of friction is dominating exists in the interior of a fluid, Such a case occurs, among others, when two layers of fluid with different velocities mect, for instance, in the wake behind 2 body, or when a fluid is discharged through an orifiee. We shall consider three examples of this type in the present and in the suceceding sections, and we shall return to them when considering turbulent flow. ‘As our first example we shall discuss the ease of flow in the wake of a flat plate at. zero ineidenen, Fig. 9.9. Behind the trailing eclge the two velocity profiles coalesce into one profile in the wake. Its width inercases with increasing distanec, and its mean velocity decreases. The magnitude of the depression in the velocity curve is directly connected with the drag on the body. On the whole, however, as we shall sce later, the velocity profile in the wake, at a large distance from the bouly, is independent of the shape of the body, execpt for a scale factor. On the other hand the velocity profile very close to the body is, evidently, determined by the boundary layer on the body, and its shape depends on wlicther or not tlh flow has separated. The momentum equation can be uscd to caloulate the drag from the velocity profile in the wake. For this purpose we draw a rectangular control surface AA, B, B, contra surface mptintcutie Fig. 9.9. Application of the momen- tun equation in the ealculation of the drag on a flat plate at zero incidence from the velocity profile in the wake 176 LX, Exact solutions of the steady-state boundary-layer equations as shown in Fig. 9.9. The boundary A,Bi, parallel to the plate, is placed at such a distance from the body that it lies everywhere in the region of undisturbed velocity, Uo. Furthermore, the pressure is eonstant. over the whole of the control surface, 80 that pressure forees do not contribute to the momentum. When calculating the fhix of momentum across the control surface it is necessary to remember that, owing to continuity, {uid inust leave trough the boundary A, B,; the quantity of fit leaving through A,B, is equal to the difference between that ontering throngl: A\A and leaving throngh B,B. ‘The boundary AB contributes no term to the momentum in the z-direstion because, owing to symmetry, the transverse velocity vanishes along it. The momentum balance is given in tabular form on the next page, and in it the convention is followed that inflowing masses are considered positive, and outgoing masses are taken to be negative. The width of the plate is denoted by b. The. total fix of momentum is equal to the drag D on a flat plate wetted on one side. Thus we have D=be f u(Uco—u) dy. (9.26) y=0 Integration may be performed from y = 0 lo y = co instead of to y = h, because for y > h the integrand in eqn. (9.26) vanishes. ence the drag on a plate watted on both sides becomes 400 2D=be f u(Veo—u) dy. (9.27) This cquation applics to any aymunetrical cylindrical body and not only to a flat plate. Et is to be remembered that in the more general case the integral over the profile in the wake inust be taken at a sufficiently distant section, and one across which the static pressure has its undisturbed value. Since near a plate there are no pressure differences cither in the longitudinal or in the transverse direction, eqn. (9.27) applies to any distance behind the plate. Furthermore, eqn. (9.27) may be appnea to any section x of the boundary layer, when it gives the drag on the portion of the plate between the leading edge and that section. The physical meaning of the integral in eqn. (9.26) or (9.27) is that it represents the loss of momentum due to friction. It is identical with the integral in eqn. (8.31) which defined the momentum thickness 5g, 80 that eqn. (9.26) can be given the alternative form D =b QU cc? by. (9.28) We shall now proceed to calculate the velocity profile in the wake, in particular, at a large distance x behind the trailing edge of the flat plate. The caleulation must be performed in two steps: I, Through an expansion in the downstream direction front the leading to the trailing edge, i.e. by a calculation which involves the con- tinnation of the Blasius profile on the plate near the trailing edge, and 2. Through an expansion in the upstream direction. ‘The latter is a kind of asymptoti¢ integration for a large distance behind the plate and is valid irrespective of the shape of the hotly. H will be necessary here to make the assumption that the velocity difference in the wake (2,9) = Veo u(x, y) (9.29) ©. Flow in the wake of flat plate at zero incidence 7 Cross-section Rate of flow Momentum in direction 2 AB 0 0 a » AA, bf Vedy eb f oot dy é é A , A BB, —b {vey —ob futdy é A a A,B, —b {Yow ay eb f Veco (Uco—u) dy é = Control surface & Rate of flow = 0 EY Momentum flux = Drag ia small compared with Uso, a0 that. quadratic and higher torms in a may be neglected. The procedure makes use of a method of continning © known solution. ‘The ealon- lation starts with the profile at tho trailing edge, calculated with the aid of Blasius's incthod, and we shall refrain from further discussing it here. ‘The asymptotic ox- pansion in the upstream direction was calculated by W. Tollinien [69]. Sinee it. is typical for problems of flow in the wake, and sinoe we shall make nse of it in the more important. turbulent case, we propose to devote seine time to au account of it, As the pressure ternt is equal to zero, the bomdary-layer equation (9.2)combined with equ. (9.29) gives au, atu, Yoo ie =" ye (9.30) whore the quadratic terms iv 1, and », have been omitted. ‘The boundary conditions are: 8% 9: y 00: 0: gf =07 yoo: y= 0. ‘The partial differential equation can, here too, be transformed into an ordinary differential equation by a suitable transformation. Similarly to the assumption (7.24) in Blasina’s method for the flat plate we put Ves YV oz? aud, in addition, we assume that u, is of the form 1" = Uae (7) tam. n where Lis the length of the plate, Fig. 9.9. The power —- } for xin equ. (9.31) is justified on the ground that. the momentumn integral whicl gives the drag on tho plate in eqn. (9.27) must be independent of =. 178 IX. Exact solutions of the steady-state boundary-layer equations Hence, omitting quadratic terms in w,, the drag ow a plate wetted on both sides, as given in equ. (9.27), is transformed to too 2D = be Um fu dy. y=-co Substituting eqn. (9.31) we obtain he 2D =b0U.2C Ve [avn dy. (9.32) Introduoing, further, the assimnption (9.31) into (9.30), and dividiug through by C Ug? + (xl)? at, we obtain the following differential equation for g (1): gang bsg =0 (9.33) with the houndary conditions g=0 at 7=0 and g=0 at y=0o. Integrating ouce, we have gting=0, whoro the constant of integration vanishes on account of the houndary coudition at 1 = 0. Repeated integration gives the solution g =exp(—4 7). (9.34) Here the constant. of integration appears in the form of 2 coefficient and can be made equal to unity without loss of generality, as the velocity distribution function u, from oqn. (9.31) still contains a free coefficient C. This constant ( is determined from the condition that the drag calculated from the loss of momentum, eqn. (9.32), mnat. be equal to that on the plate, eqn. (7.33). First, we notice that too 0 [onan = f exp (—] dy =2 yx, s0 that, from eqn. (9.32) we have 2D=2VnCbo vat Fe . On the other hand, from eqn. (7.33) we can weil wetted on both sides in the fori: i » down the skin friction on a plate =139 2 y/et 2 D = 1-328 b QUeo Viz. Toner 2. fr ~ 1-328 and C —- 0-664/)/7, and the final solution for the velocity £. The two-dimensional laminar jet 179 difference in the wake of a flat plate at zero incidence becomes ‘The velocity distribution given hy this asyinptotie equation ia represented in Fig, 9.10. It is remarkable that the velocity distribution is identical with Ganss’s error-distri- bution function. As assumed at the beginning, eqn. (9.35) is valid only at great. distances from the plate. W. Tollmicn verified that it may be used at about 2 ~ Fig. 9.11 ooutains a plot from whieh the whole velo field can be inferred. ‘The flow in the wake of a plate as well as in that behind any other body is, in most cases, turbulent. Even in the ease of small Reynolds numbers, say R, < 10°, when the boundary layer remains laminar as far as the trailing edge, the flow in the wake still becomes turbulent, because the velocity profiles in the wake, all of which possess a point of inflexion, are extremely unstable. In other words, even with comparatively small Reynolds numbers the wake becomes turbulent. Turbulent wakes will be discussed in Chap. XXIV. f. The two-dimensi ar jet The efflux of a jet from an orifice affords a further example of motion in the absence of solid boundaries to whieh it is possible to apply the boundary-layer theory. We propose to diseuss the two-dimensional problem so that we shall assume ayes Fig. 9.10. Asymptotic velocity distribution in the laminar wake behind a fiat plate, from eqn. (9.35) Fig. 9.11. Velocity distribution in the la- > minar wake behind a flat plate at zero incidences: 180 1X. Exact solutions of the steady-state boundary-layer equations that, the jet emerges from a loug, narrow slit and mixes with the surrounding fluid, ‘This problem was solved by II. Schlichting [56] aud W. Biokley [3]. In practice, in Uhis ease, as in Uhe previous ones, the flow becomes turbulent. We shall, however, discuss here the laminar case in some detail, since the turbulent jet, which will he considered later, can he analyzed mathematically in an identical way. The emerging jet carries with it some of the surrounding flail which was originally at rest. because of the friction developed on its periphery. The resulting pattern of streamlines is shown in Fig. 9.12, We shall adopt a system of coordinates with ils origin in the slit and with its axis of abscissac coinciding with the jet axis. Fig. 912. The lami tvo-dimensional free j The jet. spreads outwards in the downstream direction owing to the influence of friction, whoreas its velocity in the centre decreases in the same direction, For the sake of simplicity we shall assume that the slit ia infinitely small, but in order to retain a finito volume of flow as well as a finite momentum, it is neeessary to assume an infinite fluid velocity in the slit. The pressure gradient dpjda in tho x-direetion can hore, as in the previous example, be neglected, because the constant pressure in the anrronnding fluid improsses itself on the jot. Consequently, the total momentum in the a-direction, denoted by J, mnst remain constant and independent of the distance x from the orifice. Hence too J =o fwdy = const . 36) It, is possible to make a snitable assnmption regarding the velocity distribution if il. js considered that. the velocity profiles w(x,y), just as in the case of a flat plate at-zero incidence, are most. probably similar, beeanse the problem as a whole possesses no characteristic linear dinension, We ahall assi{me, therefore, thal the velocity w is a fanetion of y/b, whore 6 is the width of the jet, suitably defined, We shall also sume that. 6 is proportional to 7. Accordingly we ean write the stream function the form oi(t) £. The two-dimensional laminar jet. Is! ‘The two unknown exponents p and g will be determined from the following condi- tions: 1. The fox of momentum in the x-lirection ix independent. of x, according to cqn. (9.36). 2, The acecleration terms and the fri of magnitude, Lion term in eqn, (0.2) are of the same order This gives two equations for p and q: 2p —qg =0 and 2p -- 2% ~1 = p--3q, antl hence, Peyi and. Consequently, the assumptions for the independent. variable and for the stream function can be written as iy a 1= Za GAs pray), Wf suitable constant factors are inchuled. ‘Therefore, the velocity components are given by the following expressions Li, zeia ()s 3 a2 YY 2nf'), (9.37) Introducing these valnes into the differential equation (9.2), and equating the pressure term to zero, we obtain the following differential equation for tho stream fanetion f (2): PAT afr =0 (9.38) with the boundary conditions v = 0 and Anfiy = 0 at y =O, and u = 0 at y - = 70. ‘This 0. (9.39) n=O: f= ‘The sobition of equ. (9.38) is unexpectedly simple. Dulegrabing once we hove frp =o ‘The constant of integration is zero beeanse of the boundary conditions at y — 0, and the resulting differential equation of the sccond order could be integrated immediately if the first term contained the factor 2. ‘This can be achieved by the following. transformation : Exan; f= 2a PEE), | | . . where a is a free constant, to be determined later. Thos the above equation transforms into Fv 2 FF 0 (9.40) 182 IX. Exact solutions of he steady-state boundary-layer equations and the dash now denotes differentiation with respect to €. The boundary conditions are &=0: FP=0; Esco: (9.41) ail the equation can be integrated once more to give Roy Reed, (9.42) where the constant of integration was made equal to 1. ‘This follows if we put F'(0) = 1, which is permissible without loss of generality because of the free constant ain the relation between f and F. Mquation (9.42) is a differential equation of Rigcati’s type and can be integrated in closed terms. We obtain ee Jnverting this eqnation we obtain 1 —exp(—2é) Festa =| oP: (9.43) Since, further, dF/dé = 1 — tanh? , the velocity distribution can be dodnecd from eqn, (9.37) and is ua Pata VA (] tanh? &). (9.44) The velocity distribution from eqn. (9.37) is seen plotted in Fig. 9.13. IL now remains to determine the const a, and this can be done with the aid of condition (9.36) which states that the momentam in the x-dircetion is constant. Combining eqns. (9.44) and (9.36) we obtain J= foarte { (1—tanb? 2)? dé = 18 pat y'?, (9.45) a We shall assume that the flux of momentum, J, for the jet is given. It is proportional to the excess in pressure with which the jot leaves the slit. Introducing the kinematic momentum J/g = K, we have from eqn. (9.45) K\k a == 0-8255 (3%) wil and cirenlar free jet fro and (11.15) respeetively. For the two- sional jot & = 0-276 K'P y/(vz)%, and for the circular jot £ = 0-244 KV? y/yx. K and K’ denote the kinematic momentam J/g 4. Parallel streams in Inminar flow 183 and, hence, for the velocity distribution uu = 0-4543 (Jha —tanh? &), y v = 05503 ( \! (2 € (1 ~ tanh? & — tanh gy, (9.46) & =0-2762 (4) fs. ‘The travsverse velocity at the boundary of the jet is Yoo = — 0-550 (A). (9.47) shoe and the volume-rate of discharge per unit height of slit becomes Q = @ f wily, or Q = 33019 (K yay? , (9.48) The volnme-rate of discharge increases in the downstream direction, becanse flnid particles are carried away with the jet owing to friction on its boundaries. It also increases with increasing momentum. The corresponcling rotationally symmetrical caso in which the jet emerges from a small cireular orifice will be discussed in Chap. XL. The problem of the two- dimensional laminar compressible jel merging from a narrow slit was solved by 8.1. Pai [49] and M. Z. Kraywoblooki [42]. Measurements performed by B.N. Andrade [1] for the two-dimensional laminar jet confirm the preceding theoretical argument very well. ‘he jeb remains laminar up to R = 30 approximately, where the Reynolds number is referred to the efflux velocity and to the width of the slit, ‘The ease of a twoulimensional and that. of a circular turbulent, jet is discussed in Chap. XXIV. A comprehensive review of all problems involving jets can be fond in 8.1. Pai’s book [49]. g- Parallel streams in laminar flow We shall now briefly examine the layer hetweon two parallel, laminar streams which move at different velocities, and so provide a further exainple of the appli- cability of the boundary-layer equations. ‘The formulation of the problem is seen illustrated in Fig. 9.14: ‘Two initially sepurated, undisturbed, parallel strenms which move with the velocitics U, and U,, respectively, hegin to interact through friction, It is possible Lo assume that the transition from’the velocity U, to velocity Uy tales place in a narrow zone of mixing and that the transverse velocity component, v, is everywhere smali compared with the longitudinal velocity, w. Consequently, the boundary-layer equation (9.1) can be used to describe the flow in the zones I and IT, and the pressure erin may be omitted. Ina manner analogous Lo that employed for the boundary layer on a flat. plate (See. VIIe), it is possible to obtain the ordinary differential equation 184 IX. Exact solutions of the steady-state boundary-layer equations if +27" =0. (9.49) by introducing the dimensionless transverse coordinate 7 = y VU;/¥2 and the stream function py = Vv U, « f. Assuming that u/U =f‘, we are lod to the boundary conditions n=-+oo: [i=l and 9.50) na=—oco: fang. ; ‘At the plane of separation aby = 0 we must-have n=0: f=0. (9.51) heeanse y = 0 there. The solution of the differential eqration (9.49) subject to the boundary conditions (9.50) and (9.51) cannot be obtained in closed form, and a numerical method mast be employed. It is possible to obtain exact numerical solu- tions by the use of asymptotic expansions for 7 —> — co and 1 -> -} oo together with a series expansion abont 7 = 0; several such sohtions were provided by R. C. Loek [45]. The problem was first. solved by numerical integration by M. Lessen [440] starting with an asymptotic expansion for 1 -» —oo. ‘The diagram in Fig. 9.14 presents velocity profiles for A = U,/U, = 0 and 0.5. An improved numerical solntion was pnblished by W. J. Christian [10}. This special case of the interaction between a wide, homogencous jet and an adjoining mass of qniescent. air is often described by the term “plane haif-jet”. 9.14, Velocity distribution in the zone between two interacting parallel streams, after R. C, Lock [45] h, Flow in the jrlet length of a straight. channel 185, According to the investigation carried out by J. Stemhener [63], as mentioned in Chap. VU, these solutions belong to a special group of solutions of Blasius's equation (9.49). The valucs of /‘(0) and /”(0) recently calculated for various values of the velocity ratio 2 can be found in that reference, In nddition, the displacement of the zero streamline has also been ealeulated. This occurs as a result, of the cireum- stance that, the normal velocity components, » = — dy/x~ (nf’ — f) for y > | co ave not. equal ut. the two edges of the boundary Inye B.C. Lock [45] studied, in addition, the case when the two half-jets differ in their densities and viscosities, and not only in their velocities. An example of such a case is the flow of air_over a water surface. The solution now depends on the para- moter % == @2 {42/04 Jt; in addition to 2. Lock provided several exact solutions aa well aa solutions whieh were based on the momentwun integral equation, An approxinnte method was also conceived by O. 1. Potter [51]. The problem of the compressible half-jel. waa studied by D. R. Chapman [7]. Compressible flows of this type play a certain part in the calenlntion of aeparated, free shear layers in wakes [8, 13]. h, Flow in the inlet length of a straight channel As a further example of two-dimensional flow ix (ke houndary layer, we shall sider the case of How in the inlet length of a straight chanel with fat parallel walls. upstream distance from the inlet the velocity distribution is assumed to be uniform and para- bolic over the width of the channel, as indicated in Chap. V. We shall assnme that the velocity in the inlet acetion is uniformly distributed over ite width, 2a, and that its magnitude is Uo. Owing to viseons friction, boundary layers will be formed on hath walls, and their width will increase in the downstream direction. At the beginning, i.e. at amall distances from the inlet section, the houndary layers will grow in the same way as they wonld along a flat plate at zero incidence, The resuliing velocity profile will consist of two boundary-layer profiles on the two walls joined in the centre by a line of constant, velocity, Since the volume of flow innst be the same for every section, the decrease in the rate of flow near the walls which is due to friction most be compensaied by a eorrosponding increase noar tho axis, Thus the bonndary layer is formed under the influonce of an accelerated external flow, as distinct from the cnse of the fiat pinte, At, larger distances from the inlet section the two boundary layers gradually merge into each other, and finally tho velocity profile is transformed asymptotically into the parabolic distribution of Poineuille flow, ‘This process can be analysed mathomatically in ono of two ways. Kirst, the integration cun be performed in the downstream direction so that the boundary-layer growth is calenlated for an accelerated external stream, Secondly, it is possible to annlyse the progressive deviation of the profile from its asymiptotic parabolic’ distrtbution, mitegration cam proceed De thy upstreain direction. Having obtained both solutions, say in the form of series expansions, we can retain a sufficient number of terms in either of them and join the two solutions at a seetion where hoth are still applicable, Jor this way the flow for the wholo iniet longth is obtained, ‘The method which was Brat used by H, Schlichting [57], will now be outlined in briof. We assume asystom of coordinates whose axis of abscissae coincides with that of the ehan- nel, Fig. 9.15. Vor the expansion in the upstream direction we shall measnre the oreinate y from the eentre-lino of the channel, whereas for the expansion jn the dawnstream direction the or- dinate y/ will be measmred from one of the walls, ‘The inlet velocity will be denoted by Uy, and that in the central stream by U(r). We begin by writing the equation of continuity fury = U,a. (9,62) 186 IX, Exact, solutions of the steady-atate boundary-layer equations 0 02 04 06 08 10 12 14 16 uw Uo ig-9.15. Velocity distribution for laminar flow in the inlet section of a channel Introducing the displacement thickuess 5, from cq, (8.30) we can write [Ww dy =U 6,, a and with the aid of eqn. (9.52) wa can write | U(2) =U, —* U, [rede (2 (9.53) "a5 a a corde i Near the inlet scetion the boundary layer develops in the same way as on a flat plate at zero uve in unaecclerated How, so that from eqn. (7.37) we have 4, AL az =12e= Kye, 7 where =)/ (9.54) cm Vat, ig the characteristic dimensionless inlet length. Eqnation (9.53) can also be written as Ula) = Ul + Kyet Kyte. J (9.55) with Ky = 1-72, In this manner the velocity outside the houndary layer has been developed in powers of ¥7. ‘The value of K, is known from Blasius's solution for the flat plate, but the remaining cocfficients Ky, K,,... are mknown, as they depend on the boundary layer which has not yet heen determined. In the series cxpansion from the trpstream direction we assume w to(y) is the parabolic velocity clistribution, i.e. u9(y) = # Up(1 —y2/a?), velocity whose higher orders may he neglected in the first approximati igure 9.15 gives an indication of the change in the velocity profile over the inlet length. It is seen that the parabolic profile is formed at about y a/a? Uy = 0-16, s0 tat the actual inlet length is Ug = 0-1Ga(Uga/r) = 0-04(2a)-R where R denotes the Roynolkls number referred to the-width of the channel, For example at R = 2000 to 5000 the inlet length extends over 80 to 200 channel widths, Consequently, the flow docs not become fully developed at all if the chanuel is short or if the Reynolds number is comparatively large. woly) — w (2,9), where nd x’ is an additional n. i, The mothod of finite differences (87 An approxiinate method of ealenlation for the (wo«limensional case which is based on the momentim equation (see Chap. X), 08 well as merous experimental resulls which reach into the tirbnient rogion, have been’ reported in two papers by H. Hahnemann and L. Ehret [36] and [37]. The flow in the entrance of a pipe has been studio‘! by L. Schiller (55). The details of a calculation which develops the solution from the downstream direction upwards can be found in [67]. The problem of the development of the flow pattern in the inlet length of a channel was exainined critically by M. Van Dyke [71] when he formulated his secoud.- order theory, see Secs. VII and [Xm, He draw attention to the fact that the salution displayed in Fig. 9.16 represents n first-order solution and is valid anly for very hurge Reynolds numbers. For this reason it is found to show certain deviations al low Reynolds numbers from the corresponding numerical solutian of the full Navier- Stokes equations. i. The method of finite differences} Modert methods (digital computers). In recent years a large number of numerical methods has been developed for solving the Inininar boundary-layer equations. ‘These methods fall mainly under the healing of implicit finite-difference procedures and represent a development of a numerical procedure first formulated by Fluegge- Lotz, and Blottner [21]. The methods referred to are accurate and fast but require access to a digital computer, The choice of method in a given case depends on the nature of the problem considered, but is also a matter of personal preference. For a review of existing methods the reader is referred to a survey article by Blottner [5]. The method proposed herein is chosen for its simplicity and its wide range of possible applications; it differs from the carly methods in that transformed (similar- ity-type) variables are used and the step sizes are allowed to vary in the streamwise and normal direction. Some of the advantages of using transformed variables ar (a) the growth in the domain of caleulation associated with the inereasing houndary- lnyer thickness is largely eliminated; (b) the boundary-layer profiles are smoother and vary more slowly in the transformed plane allowing larger step sizes to he used; and (c) the finite-difference formulation becomes virtually identical for compressible and incompressible plane and axially syminetric boundary-layer flows. The use of variable step size in the normal direction makes it possible to calculate turbulent ns well as laminar flowa with only minor changes in formulati Special classes of laminar flows characterized by boundary layers with different length seales (c. g. large blowing rates), can also he handled with greater accuracy. The bomdary-layer equations considered are a — (rt |- —— (rt q bg (1H) ba (rv) (9.56) eu aul dp , @ ec) au wa beggar tall +e) ael eon + [au indebted to Professor T. K. Fanneloep of the Institute of Technology in Trondheim, who kindly provided me with the following presentation. 188 IX. Hxact solutions of the steady-state boundary-layer equations flow) or j= 1 (flow with axial symmetry). The boundary con- ditions are % Oat y = O and w= Ula) at y = 4. Por torbulent flows wand 9 are the appropriate mean vel« and F; represeiits a snitably defined eddy viscos- , see for instance A. M. 0. [61]. For laminar flows «= 0. he Wansformation of eqns. one and (9.57) to dimensionless variables incorporates both the Blasiua and the Mangler transformation (see also IT. Gocrtler (33, 34]) and is defined as follows: where j = 0 (pla » z = rf U(x) dx, (9.58) é z we 9.58) y= yun / (2 foerae ) ¢ é The continuity equation is satisfied by the streain function ae a (9.59) r or ‘The corresponding d nless form, {(E.), of the stream function expressed in terms of the independent variables E.1 i detined by pla) = V2ENE n). (9.60) It can now be shown that {(E, 9) satistiés the partial differential equation (N fund + f fan + BL — $8) = 2 E (ty Ing — fe fan) (9.61) Hore N14 dor, and ey is tho eddy viscosity from eqn. (19.2). The subseripts denote partial differen tiation, and the quantity wen f . BE) = 2a fuera (9.62) é is the only one determined by the flow, ‘The boundary conditions for f are n= 0; [=0; f,=Oandy= 00; fy = (9.63) Finite-difference equations of second order can be solved (by matrix inversion routines) mnch more efficiently than third (or higher) order equations. It is of inter- cst to reduce equations (9.61) to second order, To this ond the variable F = f, is introduced and eqn. (9.61) i# rewritten as IN Phy bf By b BO — Py DEL Ry —fe Fy). (9.64) This equation now contains two unknown functions, { and F, but these are related by the simple expression ” HE n) =f FE nd + f(G, 0). (9.65) i, The method of finite differences 189 In the absence of suction or blowing the boundary conditions are F(E,0) = {(E,0) = 0, F(E, c0) Near the leading edge of a eusped body and in the stagnation region of a bhint. body, equations (9.64) and (9.05) reduce to true similarity forin. ‘The corresponding similar solntions can be awed, therefore, as initial values for the slep-by-atep finite. dillerence method. The method presented here solves the partial differential cquation (9.64) nd (with small modifications) also the associated similar equations required as initial values. The method is thus solf-starting and requires no additional input. I. (9.66) ference quotients: The domain of calculation in the (é, 7)-plane can he represented by a seini-infinite strip bounded by the wall 7 = 0, the cilge of the boundary layer 7 = ne, with ye suitably defined, and the initial line € = & where the sohition is presuined to be known. This strip is completely covered by a grid with lines drawn parallel to Uh 7 coordinates as illustrated in Fig. 9.16. The step AE represents the distance between two suceessive grid lines £ = constant; it is presumed to be sinall bu otherwise unspecified. The eorresporuling step sizes Ax int the -direction are apeeified to vary in geometric progression. The ratio between two successive grid lines, 1x and nat, is denoted by K = | 4 k where || varies front 0 to 0-06 in typical case: Each nodal point is identified by a double index m. x which defines i Em, Nn according to and Em = bo + z Ab, ome = A or 11), (9.67) jon In writing the fnite-difference quotients it is convement to introduce the mean of two successive An-values Ann Sgn + Ann): (9.68) In the step-by-step calcnlations the solution is considered known at &,_ and all preceding grid lines, and the variables #* aud f are sought at Ema. Corttrnl-ditte obtained by expanding Frit. ner and Fist, na vespectively. nee Approximations to the derivatives Fy and My at Emir ina Taylor ser’ 9.10. Variable step size fitite-difference grid for the calentation of laminar and turbulent homudary layers x known vanes, unknown emues 190 IX, Exact solutions of the steady-state houndary-Inyer equations centered at. (m--1, n). The two expressions are thereupon combined in such a way that terms of order Ar? are eliminated. ‘The corresponding difference quotients can ho given the form (index m +] omitted): Tie ype Ua Pel La Put a Paes} 1 O (Age Rn), (9.60) Ayn where y= KK, bg =(K2— NI, Ig = — K; OF n Pi D ?, ?, “iit = Fae {Pa Pao — Ps Pat Po Pasi} +O (An®, kAn), (9.70) where Py= "Ye (14K), Po=ly, Pa=2Pi Pe, Pa=l. Equations (9.69) and (9.70) reduce to the standard form for central differences when K=1, For the E-dorivatives in equation (9.64) a simple backward difference formula is used "1 O (AE). (9.71) az The larger truncation crror which appears her is balanced by the iterative seheine proposed for solving the difference equation, ‘The non-linear terms in equation (9.64) have to be replaced by lincarized difference quotients. ‘The terms {'y and FF, may serve as examples and they are written as FPe= FUP )m vin and [(Py)m+ inns (9.72) where (Fp)miiin and (Fa)nitin ave given by (9.71) and (9.69), respectively. The unknown cocfficients FE and ff are set. equal to the known values Fyyy2 and fmm in the first iteration ( 0) and later updated by Fi, fi(i = 1, 2,3,...) in the second and further successive iterations. Uxperienes suggests that the tern F2 should be approximated by Pisin = 2 Fm Pain — Finn (9.73) Tho linearized finite-difference quotients given above are substituted into the differential equation (9.64) and the result. is nvultiplied through by Aé to give a differenee equation. ‘This is written as follows AnPmsin-1 } BuPmotjn cb Cn Pmainer = Day (9.74) where Toll} (Nil E2EUp. (9.75) Haale "Py Ni | > Le [fi -b (Ny) + 2 & (fp)! La] — Aik 2 Ann (9.75) — 2 Pinu (AEB -F E), i. The method of finite differences: 191 AE AE Ca 7a ™ Py Po + Dan Ty [fi + (Na) + 2 E (Ip! (9.750) Dn = —~B (1 — FR) ~2E Pg (9.75 d) In equations (9.78), € and B are evalunted at (m 1), and only the variables swith superscript i are updated through snecessive iterations. To speed-up the iterati process the torms (f,)! can be kept constant (equal to the value at the station) until initial convergence is achieved. Method of solution: Equations (9.74) represent a sct of N—I simultancous ulze- braic equations for the unknown Finyisn (m = 2,3,...,N). At cach level » three unknown quantities appear, namely Frits at, Fmiten and Prats nit, but since Fria and Fmi.w are known from the boundary conditions, the total number of cquations equals the number of unknowns. ‘The set of algebraic equations can he wrilten in so-called threc-diagonal matrix form. Matrices of this type where off-dingonal clements vanish outside the three-diagonal band can be inverted by a simple and direct method well suited for digital computers. To end this equation (9.74) is rewritten in “standard form” (subscripts (m -+ 1) omitted) AnFn-t-+ Baka -+ Cn Fnst = ni 22n<¢N—1. (9.74) The hondary conditions are Fi, =0 and Fy=1, (9.76) where n == | denotes the wall and x = N the edge of the haundary layer. It is assum- ed now that a solution exists in the form Py = Bn Fava + Gn. (9.77) The boundary condition F; = 0 and the requirement that equation (9.77) shauld remain valid independently of the step size Ay leads to By =0, = 0. (9.78) A direct, consequence of equation (9.77) is that Fa-1 = En-1 Fn} Gn-1. (9.79) When the preceding expressions are substituted into cqn. (9.74), the following relation is obtained — On Dy — An@, = P . Pa An 9 BB Ay Bay ba An Byer (9.80) A comparison between equations (9.77) and (9.80) shaws that = Cn Dn = An Gna Bn TT Ay Bae O° > BT Ay Be (al) By incans of equation (9.81) and the condition (9.78), it becomes passible to compute } Fora justification sce R. D. Richt moyor {53}. 192 1X. Exact solutions of the steady-state boundary-layer equations Fy and Gp for successive values of n starting with n = 2 for all grid points between the wall and the edge of the hoandary layer. Sinee Fas; for n = N—1 is known from equation (9.76), it hecomes possible to evaluate all unknowns Fy, by means of equation (9.77) while traversing the boundary layer from the edge to the wall through decreasing values of 2, i.e. form = N—I, N—2, ...,2.'This completes the calculation of Fy (> Pyrite) in one iteration, Once Fy jton hws been determined, the corresponding solution for fmsiya can be found by direct numerical integration of eqnation (9.65), ‘The trapezoidal rule suffices for this purpose. The caloulated values Fnsisn and fnsi,n are used to deterntine new and iniproved values of the coefficients An, Bn, Ca, which in turn leads to new and improved values of Pravin and fmirn. The process is terminated when the results of two snecessive iterations agree to within a specified tolerance, typically of order 10-5. The conver- gence is naually rapid, three to four iterations being adequate in most cases with step sizes Ar in the range 0-01 to 0-05 In certain: problems it becomes necessary to allow for bonndary-layer growth hy inereasing N (or ye) as the calenlations proceed downstream. The boundary- layer edge is defined by the requirement that the difference Fy—Iy-1 should be $s than a specificd value, typically of order 10-4, The growth, in terms of the pre- sent variables, is usually very modest even for cases involving separation. A variable of primary interest in the calenlation is the stress at the wall; its value can be determined with good accuracy froin the five point formula oF : 7 7 7 », (SE) =e ere ba Pot Prat Us 3}, (9.82) where Ty = —(L4 K + K? 4K), Ty= ra(l + K 4K) i irk’ (9.83) Ty= Initial values: When using tabulated sinilar solutions as starting values, ex- tensive interpolation is required whenever variable step sizes Ayn are used. It is more convenient and efficient. also to generate the similarity solution by finite differences through successive iterations. The equation to be solved is obtained from equation (9.64). and can be written in linearized form as Ne PY fa-a-b Nia) Fit BU — Fina Pd) (9.84) with ' , ! f= f Prd. (9.85) a valuated and ially by The indices (i, *---1) indicate the iteration for which the variable (yY denotes d/dy. Variables with index i—1 are considered known ( i, The method of finite differences 193, guessing a solution which satisfies the boundary conditions), whereas those with index 7 are to be ford in the i-th or current iteration. The difference quotients (9.69) and (9.70) are now substituted into equation (9.84). ‘The result is a difference equation which can be written in the standard form of equation (9.74), with coeff- jen ts a Ann (fi~ An News Py Pat (9.86.0) By = — Mi PyPat ba Am Vi + Nin) Ang ph Fir. (9.86) Cu = Nia Ps Pat 5 la An ea + Nie), r) Dy = — Ti, B. (9.860) A lincar variation in 7 suffices as an initial guess, Fo, and the corresponding value of / is determined from equation (9.85). The coefficienta 4x, Bn, Cn and Dy are cal- culated noxt and the corresponding valnes of By and Gy are determined across the boundary Inyer. The recurrence relation (9.77) and the boundary conditions (9.78) are then used to determine the new iterate, Fi, across the boundary layer. ’ process is repeated until the difference between successive iterates becomes smaller than the specified tolerance. The number of iterations required is typically of order 8 to 12. The method is simpler than the usual “shooting” method used for two-point boundary-value problema and it converges in many cases where the lalter method fails, for instance for very large. blowing rates. Applications: The finite-difference method presented here is intended as a prac- tical engincering tool, Greater accuracy can he achieved with a more elaborate pro- cedure, but this in turn leads to greater complexity in formulation and programming and to an increased clemand for computer time and capacity. ‘The computing time and accuracy depend for all difference methods on the step size nsed in the calen- lations. It is of interest to examine the accuracy of the present method in a few cases for which very accurate solutions are known. The cases considered are Howarth’s linearly retarded flow (cf. See. IXd) and the circular cylinder with a pressnre distri bntion according to potential theory and according to the experinents of Hiemenz (cf. See, Xe), The results for a “normal” step size and a “sinall” step size are tabulated below, From the calenlated results only the location of the separation points are shown, Cane Cousidered — [Present results | Exact a Linearly retarded flow | (1) ae* = 0:5227 ast = 01199 (Howarth) : (2) ae* = 0-1210 or 2.* = 0-1198 (Leigh)[44] or 2%* = 0-1203 (Schoenanor) Cirealar cylinder (I) be = 106-13° $e = 104-5" (Schoctianen) (Potential flow) (2) 6, = 105-01° (cf. See. Xe) Cirenlar eylinder (1) $e = 80-0°C (daife nnd Smith) Hp (Iicmenz press. data) | (2) $e (interpolnted) (1) “Normal” stop size: AE = 0-01, Ay = 0-05; (2) “Small” atop sizes AE = 0001, Ay == 0.025 194 IX. Exact solutions of the steady-state boundary-layer equations ‘The computing time with the “normal” atep size is typically 5 to 10 seconds on the UNLVAC 1108 compnter, The accuracy with the small step size is seen to be better but at the expense of a twenty-fold increase in computer time. For euginecring calen- lations the coarser grid should suffice; it requires running times of the order 10 seconds in case of practical intercat such as the laminar honndary layer of m1 acvo- foil. Improved economy can be achieved hy varying the step size as the calculation proceeds, that is using the fine mesh only in the critical region near separation. A summary account of numerical methods in fluid mechanies is given in the Ieeture notes of Sinolderen [65]. j. Boundary layer of second order} ‘The second-order equations, eqns. (7.52) and (7.53) for flow in a bonndary layer were derived in Sec. VITF. This system of linear partial differential equations can he solved if the first-order solutions w(x, N) and v; (x, N) are known, and if the fune tions K («), Uz(x, 0) and P2(x, 0) are suitably preseribed. It follows that the calculation of a second-order boundary layer on a given body in a stream requires that the following steps should be taken: (2) Calculation of the potential flow (external flow of first order) about the body with the boundary conditions V1 (x, 0) = 0. The solution yields Uj (x, 0). (b) Calculation of the first-order boundary layer for given Ui(x, 0), that is, deter- mination of the solution of the system of cquations (7.49). In particular, from the solution 1% (2, N), v(x, N) we calculate the fimetion V2(x, 0) with the aid of eqn. (7.45). (c) Calculation of the second-order external flow for the boundary conditions V2 (zz, 0) and zero velocity at large distance from the body in accordance with eqn. (7.45). The solution provides us with Up(x, 0) and P2(x, 0). In what follows, we shall assume that these steps have already been taken. We shall concentrate on more detailed second-order calculation for several particular casos. Symmetric stagnation flow: This type of flow was analyzed in detail by M. Van Dyke (sce also Chap. VII, [7)). It is assumed that the expressions for the external flow of first and second order on a convex wall at the stagnation point (K = I at x = 0) have been found and yield U(x, 0) = Ua + € Ug w + 0 (e), (9.87) where Uy and Uy, are constants which depond on the shape of the body. According to eqn. (7.48), we make the following assumption for the inner solution: u(x, ye) = Un wf'(y) +el(WUrn ele (n¥ + Var @ Fy (m)] + Ole), (9.88) ofa, ye) = —e Va f(y) — €2 [Pe(y) ~ af) + Var Fa (™)VUr1] + Ofe?). (9.89) t 1am indebted to Professor K. Gersten for the exposition in this arction. j. Boundary layer of second order 195 Tere, the new variable is defined as 9 =VOuN = VOn yet (9.90) Snbstitnting these forma into the boundary-layer equations of first and second order, we obtain lag f +l—-ft= (9.91) PO RE 2p FEA Pe = 0 ff — 1? + 2) + 06479, (9.92) PY 4 RL —2 +t Pa 2, (9.93) with the boundary conditions n=0: [=0; f'=0; Fe=0; Fi =0,Fe=0,F,=0, Foal; F, noo: f= The firat- equation determines the first-order honndary layer whioh is iden eqn, (5.39) for stagnation flow at a flat plate. The two succeeding equations deter- mine the second-order boundary layer. The solution has been split into two parts, the partial solution due to curvature (subscript ¢) and the partial solution due to the displacement effect (subscript d). The latter is induced by the external flow of sccond order with the velocity U2(x,0) = Un, as determined in step (c) above. For Fa we obtain the following simple solution i + Faas tan. (9.95) The skin-friction coefficient follows from eqn. (7.55). Inserting the numerical values 1'(0) = 12326; F2(0) = — 1-9133; F1/(0) = 1-8489, (9.96) we find that fog =e 2 Via {1.2826 Uys — e (1.9133 VTi; — 18489 Vos)} + O68). (9.97) According to eqn. (7.54), the pressure coefficient is Cow = 1 — Uf 2? (1 — 6 (1-8805//T 11 — 2 Usy/Ur1) 1 O(e2)}. (9.98) The formulae for the pressure and skin-friction coefficient are universal. The missing numerical values of the constants Uy and U1 depend only on the shape of the body. Tn all known examples, Ugi has tuned out to be negative. This signifies that the skin-friction coefficient near the stagnation point on a convex wall decreases due to higher-order boundary-layer effects (curvature and displacement); the oppoxite in true abont the-pressure coefficient at the wall. + This equation is seen to be identical with eqn. (9.16) of Sco. [Xe if it ig noted that the coordi- nates x,y in it denote lengths, whereas here they have been referred to the characteristic length Ro (radius of curvature of the horly at the stagnation point) and made dimensionless. When comparing the velocity distribution of the external flows from eqns. (9.87) and (9.15), we have n=VUaN = yluje. 196 IX, act solutions of the steady-state boundary-layer equations: Parabola in a symmetric stream: The second-order boundary layer ona parabola in a symimnetric stream was calculated by M. Van Dyke (see also Chap. VI, |7}). In the neighbourhood of stagnation, we have Uy = 1 and Uj = —061. (9.99) In the ease of the parabola we have at onr disposal a numerical solution of the com- plote Navier-Stokes equations duc to R.'T. Davis [11f and can use it for a direct evalnation of the improvement made by the second-order theory. Figure 9.17 shows a plot of the skin-frietion coefficient from (9.97) at a stagnation point of a parabola in terms of the Reynolds number formed with the radius of eurvature at the vertox. It follows from eqn..(9.97) that 12326 — plea 04 e + O(%), (9.100) fl which is equivalent to 4.2026 12476 cf2ew J Ole?) (9.101) Curve 2 in Fig. 9.17 is a plot. of this relation, whereas Curve 1 depicts the first-order solution, Curve 3 has heen plotted with the results of R. T. Davis’s numerical solu- tion, The considerable improvement. effected by the second-order theory in the lower range of Reynolds numbers is clearly visible. In addition, the diagrams give an u ambignons indication that the sccond-order theory allows us to identify the range of validity of first-order theory. Jf an crror of up to 2% is to be tolerated, it follows that first-order theory applics at Reynolds nambers in exeess of R= 1-5 x 108, Similar comparisons based on RB. 'T. Davis's numerical sohitions reveal that the lower limit of validity for the second-order theory is at R == 100 for a 2%, toleranc Figure 9.18 givos diagrams of static pressure and skin-frietion distrimutions along the contour of a parabola at zero incidence, both evahiated with the aid of second- order theory. For purposes of comparison, the diagrains contain distributions caleu- lnted with the aid of first-order bonndary-layer theory (R -+ oo). Both pressnre distributions start with cp = 1 at the slagnation point. Frietionless flow (R > 00) gives 1 P2a* " (9.102) Cp = where a* -= «// Rg denotes the dimensionless diatance from the vertex of the parabola and measnred along the centerline; sec also Fig. 9.18. Por R = 100 we find trom eqn. (9.98) for the neighbourhood of the stagnation point that tp = L— 2 (1 --3-10.6) + Ole). (9.103) This is equivalent to (9.104) 3B *,” whore x = (2a*)'? near the stagnation point. As expected, the higher-order correc- tions deorease in the downstreain direetion, particularly also duc to the decreasing enrvature in that direction, At abont a = 2, the higher-order effects vanish to all intents and purposes. Sinilar conchisions apply to the skin-friction coefficient which, however, displays at the stagnation point the largest second-order correction. - Boundary Inyer of second order 197 Fig. 9.17, Local skin-friction coefficient in the neighbourhood of a stagnation point of a parabola in termd of the Reynolds mamber R= Un R/y (1) Firstovder houndary-layer theory,R —* co: (2) Seemu-order tomdary-tayer theory, enn, (8.101), ner, Geraten (1@f of Caen. VU) (3) Numetieal eolatioh oF Ue Navier-Stokes equa: Uous after #,"P, Davis (111 (4) R= 0, Stexestan ow n 1 1 1 L n 0° 0? "0" an Re Yah (1) ep for nonviseous flow, R—> co, equ. (9.102) (2) ep for R = 100; equ. (9.104) (3) 6, R= 3-486 #4" ;alaguation polut; R-» co! eqn. (9.101); ©= 0 (4) €7 RY¥= 9-63 29!/:ataguation point; R—> co; (9.101); ©= 07 fat pinte ar 7 0 00 wx Fig. 0. 18.0) Static pressure distribution and b) distribution of shearing stress around the eontour of a parabola at. zero incidence. The curves for R = 100 correspond to second-order boundar Uheory; the curves R -> co correspond to first-order theory velayer Whereas the pressure coefficient is inereased by higher-order effects, the skin. frietion is reduced thereby. It follows that the pressure drag of a parabola increases, whereas the skin frietion decreases for Reynolds numbers decreasing from R- > oo. For the pressure drag of parabola of width b (ex [11] that nsive of bage drag), we find D cp = 2 40 U% Rob n+ GLE + Ole2). (9.105) Thus, at R = 100 the pressure drag exceeds that in an inyiseid flow hy 16%. ‘The fact that the pressure drag inercases as a reanlt of the operation of second- order boundary-layer effects points to the possibility that such drag should appear in the framework of a second-order theory also fora flat plate at zere incidence, as already intimated in See. VILE. 198 IX. Exact solutions of the steady-state boundary-layer equations Other shapes: Sccond-order effects far half-bodies have been investigated by L. Devan |12]. The results are sintilar to those for the parabola. 'The coefficients for eqns. (9.97) and (9.98) are Uy = 15; Un = — 0-62. Farther solutions of the boundary-layer equations (7.52) and (7.63) of second order are availuble, ax might have heen expected, for cases which lead to self-sintilar solu- tions in first order, Sec. VIIIb. In the case of flows whose first-order external flows are of the form Uy (x, 0) ~ 2” the second-order theory also Icads to self-similar solutions if K(x) S et@-DP2; Usher, 0) Sa, (9.106) Further details concerning the effects of second order can be found in Chap. VIL as well as in VIIT [Ga], [15a]. ‘The latter contain indications about second-order effects in the presence of suction, blowing, heat. transfer and compressibility. Second-order effocts acquire increasing importance for high Mach numbers and in the presence of blowing. In this connexion consult [24, 25, 47, 48, 59]. References [1] Andrade, B.N. ‘The velocity distribution in a liquid-in(o-liquid jet. The plane jet. Proc. Phys. Soc. London 51, 784-- 793 (1939). [2] Baxter, D. o-Lotz, 1.2 The solution of compressible laminar houndary layer problems by a finite dificrence method. Part I: Further discussion of the method and computation of examples. ‘Techn. Rep. 110, Div. Eng. Mech. Stanford Univ. (1957); short version: ZAMP 9h, 81 96 (1958). [3] Bickley, W.: The plane jot. Phil. Mag. Ser. 7, 23, 727-731 (1939), 3 ten in. Flitssigkeiten mit kleiner Reibung, Z. Math. w. Phys. 56, -3, Grenzse 1-37 (1908); ba transl. in NACA TM 1256. GF 15} Blottner, “inito difference methods of solution of the boundary-layer equations. AIAA ~ 205 (L970). [5a] Blottner, F. G.2 Investigation of some finite difference techniques for solving the boundary Inyer eqnations. Comp. Math. Appl. Mech. Kng. 6. 1 — 30 (1975). [6] Cobeci, 'T., and Smith, A.M. C 2 dilforence method for calenlating compressible laminar nnd turbulent boundary layers. ‘Trans. ASME, J. Basie Eng. 92, 523—535 (1970). {7} Chapman, D.R.: Laminar inixing of a compressible fluid. NACA TN 1800 (1949). [8} Chapman, 2.2: Theoretical analysis of heat transfer in regions of soparated flow, NACA ‘TN 3792 (1 19] Chen, KAI Skam prot and Libby, P.A.: Boundary layers with small departures from the Falkner. JEM 33, 243— 282 (1968), Improved numerical sol condition, JASS 28, 911—912 ( RT: Numerical solution of the Navior Stoke apressible flow past a parabola, JIM 57, 417-4 [12] Devan, L.: Second order incompressib! dimensional semi-infinite body, Ph, D. U3} Denison, M.R, and Baum, 1. AMAA A. 7, 342—349 (11963). | sand Gross, F.: Exnet similar solutions of the Advaueos in Heat Transfer Vol. 4, Academie Pre: [15}. Beans, W.L.: Laminar boundary Inyer theory. Addis London, 1968. [16] Fage, A. The airflow nround a circular cylinder in the region where the bowudary separates frow the surface, Phil. Mag. 7, 253 (1939). mn of the Blasins problem with t ce point 61). equations for symmetric laminar (1972), laminar houndary layer development on n two. hesis, Univ. of California at. Lan Angeles, 1964. : Compressible free shear layer with finite initial thickness. minar boundary layer equa- . New York, 1967, 317—446. Wesley Publishing Company, References 199 [17] Fage, A., and Falkner, V.M.: Further experiments on the low around a circular cylinder. ARE RM_ 1369 (1931). [18] Falkner, V.M.: A farther investigation of solution of boundary layer. ARC RM 1884 (1939). }19} Falkner, V.M.: Simplified caleulation of the laminar boundary layer. ARC RM 1805 (1941), [20] Fannelocp, T., and Fligge-Lotz, 14: ‘The compressible boundary layer aloug a wavesshaped I, Ing.-Arch, 33, 24--35 (1963), [21] Mliigge-Fotz, | layer flow ing displicem: Stanford Univ. Div. Eng. Mécanique 2, 307-423 (1963). [22] Fliigge-Lotz, L.: The computation of che laminar compressible boundary layer, Dep. Mecl. Eng. Sanford Univ,, Rep. It. 362-307 (1954). [23] Vrossling, N.: Verdunstung, Wirmeibergang aul Geschwindigkei(avertel dimensionaler wud rotationssywmetriecher lawinarer Crenzschichtatrdinuag, Lands. Un Arenlor. N. FP. Avd. 2, 36, No. 4 (1940); sce also NACA TM (432, [24] Gersten, K,, and Gross, JJ. F.:Mass-transfor effeots in higher order boundary Inyer solutions. The leading edge of a awept cylinder. Int, J. Heat Mass Transfer 16, 65—79 (1972). {25} Gersten, K., Gross, J.F., and Barger, G.: Die Grenzschicht, biherer Orduung an der Stan- linic eines achicbenden Zylinders mit starkem Aushlasen, 7. Plugwise. 20, 330341 (1972), [26] Goldstcin, 8.: Ou the two.dimensional steady flow of a viscous fluid behind » solid body. Proc. Roy. Soe. London A 142, 545—562 (1933). [27] Goldstein, S, (ed.): Modern developments in fluid dynamics, Vol. Z, 105. Clarendon Press, Oxford, 1938. [28] Goldstein, S.: On laminar boundary layer flow near a position of separation. Quart. J. Noch Apple Math, 1, 43-- 60 (1948). [29] Gortler, H.: Kin Diflerevzevverfahren zur Berechnung ta Arch. ia, Tiana) (1048). [30] Gértler, II.: Bivfluss ciner sclwachen Wandwolligkeit anf den Verlauf der laminaren Creu- schichten. Parts | and U, ZAMM 25/27, 233—244 (1947) and 28, 13—22 (1948). [31) Gartler, H.: Zur Approximation stationarer laminarer Grenzsehichtatromungen mit Hilfe der abgebrochenen Blasiugschen Reihe. Arch. Math. /, No. 3, 235 —240 (1949). {32] Gorter, H.: Reibungswiderstand einer schwach gewellien Kingsangestromten Platte. Arch, Math. 7, 450—453 (1949). [33] Gdrtler, He: Eine neue Reihenentwicklung fiir laminare Grenzschichten. ZAMM 32, 270— 271 (1952). 134) Gorter, Te A new series for the eateulation of steady laminar boundary layer ows. J. Math, Mech. 6, | —66 (1957). . [39] Gorter, H., and Witting, H.: Zu den Tanischen Greuzschichten. Oster, Ing.-Archiv 17, UL 122 (1957). [36] Hahnemann, I, and Ehret, L.: Der Drnckverlust der laminaren Stromung in der Anlauf- strecke von geraden, cbenen Spalten. Jb. dt. Luftfahrtforsehung 1, 21—36 (1941), (37) Nabnemann, H., and Ehret, L.: Der Strdmungawiderstand in geraden, ebenen Spalten unter Be chligung der Kinlanfverluste. Jb. dt. Luftfahrtforschnng 1, 186 --207 (1942). {38} Hartree, 1D. R.: A solution of the laminar boundary layer cquation for retarded flow. ARC RM 2426 (1949). 139} Hieutenz, K.: Die Grenzechicht, an cinent in dex gleichférntigen Flissigkeitsstrom cinge- tauchten geraden Kreiszylinder. Thesis Gottingen 1911; Ding!. Folyteclm.}. 326,321 (1911). {40} Howarth, L.: On the calculation of steady flow in the boundary layer near Uke eurface of a eylindor in a atroam, ARE RM 1632 (1935). [41] Howarth, 1. On the salution of the ta London A. 164, 547 ~-579 (1938). [42] Jaffe, N.A,, and Sinith, A.M.O.: Calculation of laminar boundary layers by ineans of a differential-difference method. Progress in Acrospace Sciences, Vol. 12 (D. Kiichemann, ed.), Pergainon Press, 1972. 42a} Keller, H. B,; Numerical methods in boundary layer theory. Ann. Rev. Fluid Mech. (M. van Dyke. ed.) 10, 417-433 (1978), [43] Krzywoblocki, M.Z.: On steady, Inuinar two-dimensional jets in compressible vise gnscn far behind the slit, Quart, Appl Math, 7, 313 (149). Computation of the compressible Inninai ihickuess interaction using finite dille . Tech, Rep. 131 (1962). Shortoned version in slournal de inaror Grenzachichten. Ing.- ar boundary layer equations. Proc. Roy. Soc. 200 1X. Fxact solutions of the ateatly-atate houndary-Inyer equations [44] Leigh, D.C... The laminar boundary layer equation: A method of solution by means of an automatic compnter. Proc, Cambr. Phil. Soe. 6/, 320-332 (1955). [44a] Lesson, M.: On the stability of the laminar free boundary layer between parallel streams. NACA Kop. 979 (1950); see also Sc.D. Thesis, MIT (1948). [45] Lock, R.C.: ‘The velocity distribution in the laminar boundary Inyer between parallel streains, Quart. J. Mech. Appl. Math. 4, 42~63 (1951). [46] Mills, R.H1.: A no(e on some accelerated boundary layer velocity profiles. JAS 5, 325 (1938). [47] Papenfuss, .D.: Higher-order solutions for the incompressible, three-dimensional hound. ary-layer flow at the staguation point of 2 general body. Archives of Mechanics (Warsaw) 26, 450 478 (1974). [48] Papen! Mass-transfer effecis on the threc-dimensional second order boundary: layer flow al, the alagnation point of blunt bodies. Mech. Res. Comm. J, 285 - 290 (1974). [49] Pai, S.1.: Fluid dynaunies of jets. D. Van Nostrand Company, New York, 1054. 150] Pohthansen, K.t Zur nihernagsweisen Integration der Dilferentialgleichnng cer schicht. ZAMM 1, 252-- 268 (1921). [51] Potter, 0.U.: Laminne bonndary layers at the interface of ¢ Mech, Appl. Math, 10, 302 (1957). 152] Reoves, B.L.., and Kippenhan, ©. J.: A particular class of similar so of motion and cnergy of a viscous fluid. JASS 29. 38--47 (1962). 153] Richtmeyer, 1. D.: Difference methods for initial value problems. Interscience, New York, surrent, parallel streams. ions of the o¢atis L., and Linke, W.: Druck. und Roibnngswid Zahlen 500 bis 40000. ZEM 24, 193--198 (1933). + Die Bntwickhu hwindigkoitsvertoiling (im Kreisrohr) und . LAMM. 2, 96-106 (1922). re Sirahleuausbreitang. ZAMM 13, 260-203 (19383). [57] Schlichting, Il. Laminare Knnalcinlaufstromung. ZAMM 14, 368--373 (1934). [57a] Schlichting, Il. Grenzschielttheorie, Wagl. transl. by Kestin, J.: Boundary-layer theory. 6th ed., McGraw-Hill, Now York, 1968. 168] Schroeder, K.: Bin cinfaches muncrischos Verfaliren zur Borechnung der laminaren Grenz- seliicht. ¥i3 1741 (1943); later expanded aud reprinted in Math, Nackr. 4, 439~-467 (1951). [59] Schultz-Grunow, F., and Henseler, I1.: Ahutiche Grenzachichtlisungen zweiter Orduung fair Strommngs- had ‘Temperaturgrenzschichton an longitudiaal gekriimmten Wainden init Arenzschichtbecinflussung. Wirme- und Stolliibertragnng 1, 214— 219 (1968). [60] Smith, A.M.O., and Clutter, D.W.: Solntion of the incompressible boundary Inyer equa- tions. AIA J. 1, 2062--2071 (1903). [61] Smith, A.M.0., and Cebeci, T.: Numerical solution of the t tions. MeDonnell-Dongins Rep. No. DAC 33735 (1967). [63] Stoinheucr. J.: Similar solutions for the laminar wall jot ina decolerating onter flow. ALAA J. 6, 2198-2200 (1968). 164] Stowartson, K.: Purther solutions of the Kalkucr-Skan equation. Proc. Cambr. Phil, Soe. 50, 454 465 (1954). [65] Smoideren, 6: Numerical methods in thiil dynamics. AGARD Lecture Ser. No. 48 (1072). [68} Tani, L.: On the solution of the laminar boundary lnyer equations. J. Phys. Soc. Japan 4, 149-154 (1940). Seo also: Fifty years of boundary-layer researeh (W. Tollmien and JL. Gortler, cd), Braauschweig, 193-200 (1955). 167] Thom. A.: The laminar bonndary tnycr af the front part ofa cylinder, ARC RM L176 (1928); see aly ARC RM 1194 (1925). 168] Tifford, ALN.: Neal, Wansfer and frictional effects in laminar be Universal series sobitions. WADE Techn. Rep. 53 - 288 (1954). {69} ‘Vollnien. W.: Grenzschichten. Handbuch der Exper. Physik #1, Part 1, 241 —287 (1981). 170] Ulrich, 3.: Die ebrne lawinare Roibungsschicht an einem Zylindor. Arch. Math, 2, 33-41 (1940). t 171] Van Dyke, M.: Bntry flow in a chaunel. JPM 44,81 172] Wilting, 11: Ober avei Ditlerenzenverfahren der 250 (1053). 173] Auovywons: luterpolation and allicd tobles. Propared by H.M. Naw ILM. Stationary Office (1956). and des Zylinders bei Reynolds. Imlent boundary-Inyer equa- dary layers, Part 4: 823 (1970). zschichtthvoric, Arch. Math, 4, 247 ical Almanae Office. CHAPTER X Approximate methods for the solu of the two-dimensional, steady boundary-layer equations Introductory remark: The exaniples of exact solutions of the boundary-layer equations which have been discussed in the preceding chapters have shown that the mathematical difficulties associated with analytic solutions for them are consider- able. In particular, the general problem involving the flow of fluid round a body of arbitrary shape cannot be completely solved with the aid of the analytical methods presonted thus far. The numerical or step-by-step methods (see Sec. IXi) allow us to solve most. problems with a tolerable amount of work if a fast digital computer ix available. For this reason, the approximate incthods for the solution of our boundary- layer equations developed in carlier times, that is before the advent of computers. do not enjoy the same importance now as they did then. Nevertheless, we propose to give here an outline of these approximate inethods, becanse they are well-suited to the generation of a quick ontline of a solution even in more complex eases; in this connexion the summary by I. Truckenbrodt [24] will be found helpful. ‘This chapter deals with approximate methods for laminar boundary layers only. Analogous methods for turbulent: boundary layers (ef. Chap, XXI) have retained their special importance up to this day. All approximate methods are integral methods which do not attempt to satisfy the boundary-layer equations for every streainlinc; imstead, the equations are satis- fied only on an average extended over the thickness of the boundary layer. All approxinate methods are based on the momentum and energy equations of boundary- layer theory known to us from Sec. Vile. All these methods can be traced to two papers, one due to Th. von Karman [7}, and the other to K. Pohlhansen [15]. Be- fore proceeding to apply the method to the general cases of two-dimensional and axially-symmetrical boundary layers with pressure gradients, we shall consider first the essential features of the method as applicd to the flat plate at zero inei- dence. This example is particularly simple in that the pressure gradient vanishes along the whole plate. Morcover, we shall have the opportunity of assessing the power of the approximate inethod, at Icast in a particnlar case, and to campare it with the exact. solution which is already known from Chap. VIL. a, Application of the momentum equation to the flow past a flat plate at zero incidence Applying the momentum equation to the fluid within the control surface shown in Fig. 10.1, we can derive the statenient that the flux of momentum through 202 X. Approximate methods for steady equations the control surface, considered fixed in space, is equal to the skin friction on the plate D(z) from the leading cdge (a =0) to the current section at x. The application of the momentum equation to this particular case has already been discussed in See. IXf. It was then found, eqn. (9.26), that the drag of a plate wetted on one side is given by D(x) = bo [ U(Uoo—w) dy, (10.1) v=o where the integral is to bo taken at section x. On the other hand the drag can be expressed as an intogral of the shearing stress ty at the wall, taken along the plate: D(x) =b [ w(x) dx. (10.2) a Upon comparing eqns. (10.1) and (10.2) we obtain (2) =ed fuw.—4 dy. (10.3) yoo This equation can be algo deduced in a purely formal way from the boundary-layer equation (7,22) by first integrating the equation of motion in the #-direction with respect to y from y = 0 to y = oo. Equation (10.3) is, finally, obtained without dif- ficulty if the velocity component v is climinated with the aid of the equation of continuity, and if it is noticed that 1.(8u/2)yao = To- Fig. 10.1. Application of the momentum equa- tion to the flow past a flat plate at zero incidence Introducing the momentum thickness, 5g, defined by eqn. (8.31), we have 45, % U2 Gt = o (10.4) ‘The momentum equation in its form (10.4) represents a particular case of the general momentum equation of boundary-layer theory as given in eqn, (8,32), being valid for the case of a flat plate at zcro incidence, Tis physical meaning expresses the fact that, the shearing stress at the wall is equal to the loss of momentum in the bonndary layer, because in the example under consideration there is no contribution from the pressnre gradient. “So far eqn. (10.4) introduced no additional assumptions, as will be the case with the approximate method, but before disenssing this matter it might be uscfi to note a relation between t and 6, which is obtained from eqn, (10.4) by introcn a, Application of the momentum equati to the flow past a Mat plato at zero incidence 203 the exact value for ty from eqn, (7.32), Putting ty/o Uo? =a V'¥[Uoo x with « = 0-332, we have Vets dz =2a 5b and hence sta z. (10.5) With reference to eqn. (10.3) or (10.4) we can now perform an approximate calonlation of the bonndary layer along a flat plate at zero incidence. The esaenco of the approximate method consists in assuming a suitable expression for the velocity distribution u(y) in the boundary layer, taking care that it satisfies the important boundary conditions for u(y), and that it contains, in addition, one free parameter, such as a suitably chosen boundary-layer thickness which is finally determined with the aid of the momentum equation (10.3). In the particular case of a flat plate at: zero incidence now being considered it is possible to take advantage of the fact that the velocity profiles are similar. Hence we put =1 (95) =o. | (10.6) where 7 = 4/6(z) is the dimensionless distance from the wall referred to the boundary- Inyer thickness. The similarity of velocity profiles is here accounted for by assuming that /() is @ function of 7 only, and contains no additional free parameter. The function { must vanish at the wall (7 = 0) and tend to the value 1 for large values of 1, in view of the boundary conditions for u. When using the approximate method, it is expedient to place the point at which this transition occurs at a finite distance froin the wall, or in other words, to assume a finite boundary-layer thickness 6(z), in spite of the fact that all exact solutions of the boundary-layer equations tend asymptotically to the potential flow associated with the particular problem ‘The boundary-layer thickness has no physical significance in this connexion, being only a quantity which it is convenient to use in the computation. Having assumed the velocity profile in eqn. (10,6), we ean now proceed to cvahiate the momentum integral (10.3), and we obtain 2 1 [u0.— way i U2 5(x) f (1 —f)dy. (10.7) v=0 n=0 The integral in eqn. (10.7) can now be'evaluated provided that n apecitic asaumplion is mace for (7). Putting 1 a, = fil — fy dy (10.8) é for short, we have [Wa — w) dy = Veo? by = 0% 6 en? v0 or by = 048. (10.9) 204 X. Approximate methods for steady equations ‘The value of the displacement thickness 5, from eqn. (8.30) will now also be caleu- lated as ib will be required later. Putting 1 a, = f(l—f)dy, (10.10) j we obtain 5, -- ad. (0d) Furthermore, the viscous shearing stress at the wall is given by t _ , (eu _ "Um Uo sar (™) 10) = B, “BE, (10.12) where By =f (0). (10.13) Introducing these values into the momentum equation (10.4), we obtain ah, * ay os” Integration from 6 =0 at x ~=0 gives the first, result for the approximate theory in the form 2. d(x) =) mE (10.14) Hence the shearing stress at the wall from eqn. (10.12) becomes X Uc Tolx) Vo n0n fe. (10.15) Finally, the total drag on a plate wetted on both sides can be written as 1 2D =2b f ryde, ic. é 2D =2bY20,f, VuolU., (10.16) and from eqns. (10.11) and (10.14) we obtain the displacement thickness 6 = “hy, (10.17) A comparison of the approximate expressions for the boundary-layer thickness, for the shearing stress at the wall, and for drag with the respective formulac of the accurate theory, eqns, (7.37), (7.31) and (7.33), shows that the use of the integral momentum equation leads in all cases to a perfectly correct formulation of the equations. Tn other words, the dependence of these quantities on the current length, x, the free-stream velocity, Uso, and the coefficient of kinematic viscosity, », is correctly deduced. Furthermore, the relation between momentum thickness and shearing stress at the wall given by eqn. (10.5) can also be deduced from the approximate caleulation, as is casily verified. The stillamknown coefficients a, %, and B, can only a. Application of the momentin equation to the flow past a flat plate at zero incidence 205 be calculated if a specific assumption regarding the velocity profile is made, i. ¢. if the fonotion f() from eqn. (10.6) is given explicitly. When writing down an expression for {(1), it is necessary to satisly certain boundary conditions for u(y), i.e, for {(n). Ab least the na-slip condition u — 0 at y = 0 and the condition of continuity when passing from the boundary-layer profilo to the potential velocity, m= Unt y+ 3, must be sntistied. urther cou ditions might include the continuity of the tangent and curvature al the point, where the tw solutions are joined, In other words, we may scck to satisfy the con- ditions dujay = 0 and 0%u/2y? = 0 at y = 0. In the case of a plate the condition that 2u/ay? = 0 at y = 0 is also of importanee, and it can be seen from eqn. (7.15) that it is satisfied by the exact: solution. Numerical examples: We now propose to test the uscfulness of the preceding approximate method with the aid of several examples. The quality of the result depends to a great extent on the assumption which is ade for the velocity tinction (10.6). In any case, as already mentioned, the funetion {(7) must vanish at 17 = 0 in view of the no- slip condition at the wall. Moreover, for large values of 1 wo must have {(7) = 1. Tf only a rough approximation is desired, the transition to the value /(17) = 1 may oceur with a discontinuons first derivative. For a better approximation, continnity in d{/dy may be postulated. Independently of the particular assumption for {(7) the quantities ‘iz nt 4 Joe oot 1%, o, ye? pw UaV Ue! ol > must he pure numbers. ‘They can be easily calenlated from eqns. (10.8) to (10.17) Use Us, | ol Table 10.1 contains results of several calenlations with alternative velocity distribution finetions. The first two finetions are illustrated with the aid of Fig. 10.2. ‘The linear fimetion satisfies only the conditions /(0) = 0 and {(1) = 1, whereas the cubic function satisfies in addition the conditions /’(1) — 0 and f”(0) = 0; finally, a fourth degree palynomial can be mae (o satisfy the additional condition {” (1) = 0. The sine function satisfies the same boundary conditions as the polynomial of fourth degree, exeept for (1) =0. Lhe polynomials of third and fourth degree and the sine-function lead to vahics of shearing stress at the wall which are in error by less than 3 per cent and may be considered entirely adequate. ‘The values of the displacement thickness 6, show acceptable agreement with the corresponding exact values. Fig. 10.2. Velocity distribution in the boundary layer on a flab plate at zero incidence Q) Linear approximation 2) Cobic approximation from Table 10.1 206 X. Approximate mothods for steady equations Tablo £0.1, Results of the caloulation of the boundary Inyer for a flat plate at zero incidence based on approximate thoory Uo] ee Uool\# | 8) a loa! f co] to y/ee «(%') OW, ufU = fin) ‘yj rx |pUcoV Uo) IV + a pfu I Sinn 6} | i | be 0-289 1155 3-00 a] 3] 3 = — 3 ae | = . . . 2] f= tn- by aso we | | 1740 0-323 L292 270 3] son = 2n— 29 + nt [EL] 3] | tree 0343 1372 255 315 | 10 - nm 4—n|lx—2] 2 4[ f(y) = sin (3 1) Balals 174th 0327 1310 2-66 5 exact —|—|— | ta 0-332 1328 259 Woo _ 219 4 / va Ucol\$ _ Uso % Vez = nto V Um o(%e')! <2 V7 It is seen that the approximate method leacls to satisfactory results in the case of a flat plate at zero incidence, and the extraordinary simplicity of the caloulation is quite remarkable, compared with the complexity of the exact solution. b, The approximate method due to Th. von Karman and K. Pohlhansen for two- dimensional flows We now propose to clevelop the approximate method of the preceding section so that it can he applied to the general problem of a two-dimensional bonndary layer with pressure gradient. ‘The methorl in its original form was first indicated by K, Pohlhausen [15]. The succeeding description of the method is based on its mare macern form as developed by HH. Holstein and 'T. Bohlen [5]. We now choose, as hefore, a system of coordinates in which x denotes the are measured along the wetted wall and where y denotes the distance from the wall. The basic equation of the mo- Mentum theory is obtained by integrating the eqnation of motion with respect to ¥ from the wall at y = 0 to a certain distance h(x) which is assmned to be ontsicle the homndary layer for all valnes of «. With this notation the momentim eqnation has the forin already given in (8.32), namely Ww dx ee. v2 4. 25, + 4) U (10.18) This equation gives an ordiuary differential equation for the hommdary-layer thickness, as was the case with the flat plate in the preceding section, provided that a suitable b. The approximate method duo to Th. von Kérmén and K. Pohthansen 207 forin is assumed for the velocity profile. This allows us to calculate the momentum thickness, the displacement thickness, anc the shearing stress at the wall. In choosing a suitable velocity finetion it is necessary to take into account the same considera- tions as hefore, namely those regarding the no-slip condition at the wall, as well as the reqnirements of continuity at the point where this solution is joined to the potential solution, Furthermore, in the presence of n pressure gradient the function must admit the existence of profiles with and without a point of inflexion corresponding to their occurrence in regions of negative or positive pressure gradients. In order to be iv a position to calculate the point of separation with the aid of the approximate method the existence of a profile with zero gradient at the wall (2u/2y)y.9 =O must also he possilile. On the other hand functions postulating similarity of velocity profiles for varions values of x may no longer he preseribed. Following K. Pohl- hansen we assume a polynomial of the fonrth degree for the velocity function in terms of (he dimensionless distance from the wall 7 = y/S(z), ic. we put Fan) san +b yt +e td yt (10.19) in the range 0 < 17 <1, whoreas for 7 > 1 we assume siinply u/U = 1. We farther demand, as before, that the boundary layer should join the potential flow at the finite distance from the wall y = 4(z). In order to detennine the four free constants, a 6, ¢, d, we shall prescribe the following four boundary conditions y=0: (10.20) y=d: usU; As seen from eqns, (7.10) to (7.12), they are all satisfied hy the exact solution. ‘These requirements are sufficient to determine the constants a, b, ¢, d, because the no-slip condition at the wall is implicit in eqn. (10.19). The first condition which is satisfied by all exact solutions, as seen from eqn. (7.15), is of particular importanee. It determines the curvature of the velocity profile near the wall and makes sure that there is no point of inflexion in the velocity profile in regions of decreasing pressure. Furthennore, regions of inereasing pressure contain points of inflexion as required by the exact solution in Chap. VIL, Figs. 7.3. and 7.4, Introdueing the dimensionless quantity _ oa A=" 4. (10.21) we obtain the following expressions for the coefficients in equ. (10.19): A A A a=2+ 4, ba}, c= 244; and hence for the velocity profile: = F(y) 4+ AG(y) =(2 9 —293 4-08) + 4 (7 —3 92-342 — 94), (10.22) 208 X. Approximate methods for steady equations where F(n) = 29 —2 9 + yt =1—(1— HF (1-0), G(n) = 4 (9 — 39? -+ 39 — nt) = 4h nL — 0). It is casily recognized that the velocity profiles expressed in terms of 9 = y/d(x) constitute a one-parameter family of curves, the dimensionless quantity A being a shape factor. The dimensionless quantity A which may also he written as _ au ap 6 A= 5 de = 7 de aps } (10.23) can be interpreted physically as the ratio of pressure forces to viscous forces. In order to obtain a quantity to which real physical significance can be ascribed, it would be necessary to replace 4 in the above definition by a linear quantity which itself possesses physical significance, such as the momentum thickness. This will be done later in this section. 12 u 10 Uv 08 06 04 02 0.2 Vig. 10.3. ‘The functions (4) and G(1) for Vig. 10.4, The one-parameter family of velo- the velocity distribution in the boundary city profiles from eqn. (10.22) layer from equs. (10.22) and (10.23) ‘The two functions F(7) and G (1) defined by eqn. (10.23), which together compose the velocity-distribution function given in eqn. (10.22), are seen plotted in Fig. 10.3. Velocity profiles for various values of A are shown in Fig. 10.4. ‘Che profile which corresponds to A -=0 is obtained when dUjda= 0, i.e. for the boundary layer with no pressure gradient (flat plate at zero invidence), or for a point where the velocity of the potential flow passes tht ongh a minimum or a maximum, In this ease the velocity profile hecomnes identical with the fourth-legree polynomial used for the flat plate in the preceding section. The profile at. scparation with (/2y), i.e. with @ => 0, occurs for A == — 12. It will he shown later that the profile at the stagnation point corresponds to A = 7-052. For A > 12 values x/U > | occur b. The approxiinate incthod due to Th. von Kérmén and K. Pohlhansen 209 in the boundary layer, but this must be excluded in steady flow. Since behind the point of separation the present calculation based, as it is, on the boundary-layer concept, loses signifieance, the shape factor is sccn to be restricted! to the range — Wx AX +12, see Fig. 10.4. Before proceeding to calculate the boundary-layer thickness 5(x) from the momentum theorem, it is now conveniant to ealeulute the momontam: thickness, 65, the cisptaccinent thickness, 5,, and the viscous shearing stress at the wall, to, with the aid of the approximate velocity profile in the same way as was done for the flat: plate at zero incidenee in the preeeding section. Thus we obtain from eqns. (8.33) and (8.31), together with eqn. (10.22), 4 = [Fin —Aamian, : $= [rot 46m] 0 FO) — AG An. 120 Computing the definite integrals with the aid of the values of (7) and G(y) from eqn. (10.23), we have A 5, —: B= (10.24) Similarly, the viscons stress at the wall, t) = Men is given by CG ya? 44 s° (10.25) In order to determine the still-umknown shape factor A(x) and, henee, the function 6(r) from eqn. (10.21), iL is now necessary to refer to the momentum equation (10.18). Multiplying by 54/7 U we can represent it in the following dimansionless form: US, oe 1) Ut tod +(24 ) y= a0 (10.26) in which the boundary-layer thickness 0 does not appear explicitly; this ciremmstance is not surprising, because it constitutes only a fortuitons quantity associated with the approximate method of calculation and has no particolar physical meaning. On the other hand eqn. (10.26) contains the really important plysieal quant viz. the displacement thickness, 5,, the momentnn thickness, J), and the shearing, stress at the wall to. It is, therefore, natural to begin with the calculation of J, from the momentum equation (10.26) and to deduce 5 from it with the aid of eqn. (10.24). Following Il. Holstein and T. Bohlen [5] it is convenient to intraiuce for this purpose a secoud shape factor 6,2 aU vo dx K= (10.27) 210 X. Approximate methods for steacly equations which is connecter with the momentum thickness in the same way as the first shape factor, A, was connected with the boundary-layer thickness, 3, in eqn. (10.21). In addition we shall pnt =| (10.28) so that, au K=25. (10.29) It is seen from eqns. (10.21), (10.27) and (10.24) that the shape factors A and K satisfy the universal relation 1 Aah At) A 10.30) a5 4 d672 . (10.5 Denoting (10.31) and a4 ) = fy(K) (10.32) for the sake of brevity, and substituting, K and Z% from eqns. (10.27) and (10.28) respectively, together with /,(K) and f,(K) from eqns. (10.31) and (10.32), we obtain, further, from the momentum equation (10.26) together with 5g d9'/v = 4dZ/dx, the relation 5 un 2 1 2 - ((K))K = fy(K). (10.33) Finally, we introdnce the additional abbreviation 2f_(K) —A K ~2Kf,(K) = F(R) (10.34) or, written ont fully, 1 116 B(K)= =2(% mal op ate) [2 ~si54 + (ais + ite) 4 + 698 34 4"), (10.35) te the relation between A and A was given in eqn. (10.30). With all these abbre- ions and snbstitntions the momentum equation (10.33) can now be rewritten in the very condensed form: (10.36) f The quantity /,_ — 34/54 is also regarded as n shape factor particular importance for the-tuebuleat, hoimdary ¢/. Chap. X11. Its value for kuninar boundary Inyers range from abont. 2-3 to 3-5, ef. He 10.2; it assumes values from about 1-3 to 2-2 in the ease of tarhtlont honudary layers. Al the point of transition JT, increases considerably, cf. Fig. 16.5. b. The approximate method due to Th. von Kérm&n and K. Pohlhausen 2u1 ‘This is a non-linear differential equation of the first order for Z = d,%/v as a function of the current length coordinate, x. The fact that the form of the function F(K) is very complex does not constitute real difficulty insofar as the solution of eqn. (10.36) is concerned, because it is a universal fuuction, i.c. one which is inde- pendent of the shape of the body and it can, therefore, be ealeulated once and for all. The functions K(A) from eqn. (10.30), as well as f,(K), f9(K), and F(K) from eqns. (10.31), (10.32), and (10.35), respectively, are given in Table 10.2. The auxiliary funetion F(X) is represented graphically in Fig. 10.6. Solution of the differential equation for momentum thickness: Conecrning the solution of eqn. (10.36) it is possible to make the following remarks: The caleulation should begin at the stagnation point x =0, where U =0 and dU/dz is finite and different, from zero, unless the body possesses there a sharp cusped edge with zoro angle. The initial slope of the integral curve d%/dz would become infinite at the upstream stagnation point were it not for the fact that F (i) vanishes there simul- tancously. Thus the function F(X) is seen to have a physically meaningful initial value. The zero of F(X) occurs for values of A for which the second bracketed term on the right-hand side of equ. (10.35) vanishes. Thus F(K) =0 for K = K, =0-0770, or for A = Ay = 7-052. ence A = 7-052 is the value of the first shape factor at the stagnation point, as already mentioned. In this manner the initial slope of the integral curve at the upstream stagnation point is seen to be of the indeterminate form § (singular point of eqn. (10.36), but its value can be easily computed by a simple process of going over to the limit. We obtain K, __ 00770 , IZ 0" By = ye = ai ((2),= — 0-0652 we (10.36.a) Here the subscript 0 refers to the upstream stagnation point. With these initial values the equation ean be conveniently integrated, e. g. by the method of isoclines. Figure 10.5 illustrates the use of this method as applied to a symmetrical aerofoil at zero incidence. The calculation begins with the values A, — 7-052 and K, = 0-0770 at the leading-edge stagnation point, and becomes completed upon reaching the point of separation with A = — 12 and K 0-1567. The velocity function U(z), together with its first derivative dU/da, is given by the potential- flow solution. The value of d?U/dx? is only required at the leading edge for the initial slope of the integral curve. The procedure usod in the computation may be summarized as follows: |. The potential flow function U(z), together with its derivative dU/da, are given in terms of the are length. 2. Integration of eqn. (10.36) gives Z(x) and the sceond shape factor K (2) so that the momentum thickness 6,(x) can be caleulated from equation (10.27), and the position of the point of separation may be found subsequently, 212 X. Approximate methods for steady equations ‘Table 10.2, Auxiliary functions for the approximate calenlation of laminar boundary layers, after Holstein and Bohlen [5] A K F(K) Ww 0-084 = 0.49658 4 0-0928 | —0-0885 13 0.0941 — 00914 12 00948 ~ 0-048 u 0-0941 —0-0912 53 5 10 0919 | —0-n800 2-260 0351 9 0-0882 = 0-04508 2-273 0347 8 0-0831 00335 2-289 0-340 78 00819 | - 040271 2-293 0338 76 0:0807 | —0-0203 2-297 0337 14 00794 | —0-0132 2:301 0335 72 00781 —0:0051 2305 0 1-052 0-070 0 2-308 0332 7 0.0767 0-0021 0.331 68 0-0752 09-0102 0-330 66 0:0737 0.0186 0-328 ot 00721 0.0274 0-326 62 0-0706 0-0363 0-324 6 0-068 0-0459 0321 5 0-0599 0.0979 0-310 4 0.0497 01579 0-297 3 0.0385 02255 0-283 2 0.0264 0:3004 0-268 1 00135 0-3820 0-252, 0 0 04698 2-554 0-235 —0 0140 0.633 2-604 0217 —0-0284 2647 0-199 0.0429 2-716 0-179 0.0575 2719 0-160 0 0720 2847 0-140 0.0862 2-921 0-120 00999 0-100 01130 3-085 0.079 01254 3-176 0.059, —0-1369 | 3276 0-039 —0-1474 3-383 0019 ' —l2 ~0-1567 17241 3500 0 13 01648 L-8i6o 3-627 —0019 ONT 1-9033 3-165 0-037 01767 1.9820 3-916 0054 b. The approximate method due to ‘Th, von Kérmén and K. Pohlhausen 213 3. The variation of the first shape factor A(2) is obtained from eqn, (10.30) and Table 10.2. 4, The displacement thickness, 6,, and the shearing stress at the wall, tp, are found from eqn. (10.31) and (10.32), respectively, together with the values in ‘Table 10.2. 5. The boundary-layer thickness (x) follows from eqn, (10.24). 6, Finally, the velocity distribution is found from eqn. (10.22). . 10.5, Example of the eal dary layer by the approximate method duc to Pohlliaisen and Holsicin.Boltlen [8]. Solution of tho differential equation (10.38) by the methods of isoclines for the synonctrieal Zhukovekii acrofoil +} O15 aban incidence angle « ~ 0, See also lig, 10,12. $+ pont of separation A, Walz, [26] pointed out that eqn. (10.36) can be reduced to a simple quadrature hy the introduction of a further approximation without any appreciable loss of accuracy, He found that the funetion F(X) can be approximated quite closely by the straight line F(K) --a—bK With a =-0-470 and 6 =6 the approximation is parti stagnation point and the poit of maximum velocity ( eqn. (10.36) reduces to arly close between the 10.6) In this manher d% dz —bK or, substitnting the original values for Z and K, (288) oo HE, dz \ » ‘This differential equation for U 4,2/v ean be integrated explicitly to [o-ae, 214 X, Approximate methods for steady equations or, using the numerical values for a and 6 given earlicr: U6, 0470 o_o [use. (10.37) sto Thus the solution of eqn. (10.36) is secn to reduce to a simple quadrature, An anal- ogous quadrature will be used in Chap, XXII for the solution of the equations of turbulent flow, Fig. 10,6, The auxiliary function F(X) for the eal- culation of laminar boundary layer by the method of Holstein and Bohlen (5} (2) ning ean. (19.98); (2) tinear approximation F(R) = 0470-—6 Ky 'S = blngnation point; Af = velocity maxtmuyn 077 0H -006 -Qos -tm -an2 0 ate Om ANG 078 « ¢. Comparison between the approximate and exact solutions 1. Flat plate at zero incidenee- It is easy to see from cqn. (10.22) that the Pohlhausen upproximation becomes equivalent to Example 3 in Table 10.1 for the case of a flat plate at zero incidence. This case can also be obtained directly from eqn, (10.36), where U(x) = Uso, U' = 0 and hence K = A = 0,80 that en. (10.36) gives dZ/dz = F(0)/Uco = 0-4698/U.o. Taking into account that Z=0 at x =0 it follows that Z—0-4698 x/Ueo, or 5, == 0-686 Vy 2/Uq, in agreement with Table 10.1, Table 10.1 contains exact and approximate values of the boundary-layer para- ineters for the purpose of comparison, It is scen that agreement is very satisfactory. 2, Two-dimensional stagnation flow. ‘The exact solution of the problem of two-dimensional stagnation flow for which U(z) = U' +x, was given in Sec. V9, The exact values of displacement thickness, momentum thickness and shearing stress at the wall, calculaicd with the aid of that theory, are given in Table 10.3. e. Comparison between the approximate and exact solutions 215 Table 10.2, Comparison of exact and approximate valies of the boundary-layer parameters for the case of two-dimensional stagnation flow Approximate method «due to 0-641 0-278 149 231 KK, Pohlhavsen exact, solutioti 0-648 0-292 1.233 221 Tn the approximate inethod we have Z, = Ko/U‘, and from eqn, (10.36) it follows that the momentum thickness is given by 6, VU']y = VK, = ¥0-0770 = 0-278 . It is seen from eqn, (10.31) that the displacement thickness is approximated by 5, VO'jv =f, (Ko) ¥ Ky =0-641 and eqn. (10.32) gives tou U «Vx/U" =12(Kolly Ky 0-332/0-278 = 1-19 for the shearing stress at the wall, The agreement between the approximate and exact values is hore also completely satisfactory, 3. Flow past a circular cylinder. A comparison of the result. of the approximate calculation for a cireular cylinder with the solution due to Hiemenz, (Sec. 1Xc) was given by K. Pohlhausen [15] in his original paper, Ic used ITicmenz’'s experimen- tal pressure distribution function for the circular cylinder and compared the results with Hiemenz's solution which takes into account only the first three terms of the Blasius series, Uicnicnz’s solution indicates that separation occurs at an angle 2-0°, whereas Pohlhansen’s approxiinate value was = 815°. However, the approximate method leads to values for the boundary-layer thickness near the point of separation which are considerably larger than the values obtained by Iliemenz, On the other hand it must be realized that such a comparison is not conclusive, because a Blasins series containing only three torms is in itself inadequate to represent, the solution near the point of separation, We now propose to give a comparison between a set. of calculations obtained with the aid of Pohlhausen’s approximate method and numcrical caleulations which have been performed with great accuracy on n digital computer programmed to solve the differential equations directly, The example choson for comparison is that of a circular cylinder in the presence of a frec-stream velocity computed from potential theory, the boundary-layer velocities having been calculated with a Blasins scrics cantaining terms up to 2x"! (See, 1X). This comparison shows thnt: the power-series method gives very high accuracy up to the immediate vie of the point of separation, However, at the point of separation itself, the series broken off at the term 2" becomes inaccurate, Figure 10.7 shows a plot of the boundary-layer parameters, displacement thicknoas, 6,, momentum thickness, 5, and wall shearing stress, tp, It is secn that the recent numerical calculations performed by W., Schoenaner [20] show somewhat different trends in the vicinity of the point of scparation as far as the variations in the displacement and mom! m thicknesses as 216 X. Approximate methods for steady equations well as in the shearing stress are concerned, and predict an carlier point of separatio: W.Schocnaucr found that the separation angle isatds = 104-5° as against ps = 109-5° obtained with the aid of the Pohlhausen approximation and ¢s = 108-8° suggested by the scrics expansion continved up to the term x. A comparison between the velocity distributions, Fig. 10.8, leads to the conclusion that there oxists almost. perfect agreoment, hetween the exact solution and the approximation in the range of angles 0 <4 <0", that is in the range of accelerated external flow. By contrast, down- stream of the pressure minimum the discrepancies increase very fast on approaching the point of separation. No general criterion regarding the admissibility of the approximation has been given so far, and it scems that it will be difficult to obtain, Judging by the ahove and similar calculations as well as by experimental results it appears, however, to be reasonably certain that Pohlhansen’s approximate method leads to very satisfactory results in regions of accelerated potential flow, Similarly, it may be a 20 - (/ iy 15 “po ~ Sehies expansion ~| ! ——-bigital computer af. | F Separron op vel | : 4 a 4 wo 0 degs ‘00 02 OB sal wa vo Vig. 10.7, Comparison of Pohlhauson’s ap- ! Fig. 10.8. Comparison of proximate solution with the exact solvtion Pohlbavsen’s approximate sa- for tho caso of a circular cylinder lution with the exact solution 4, ‘ aisptncement thickness: for the case of a circular cy: 8, = momentnin thickness; livders velocity profiles tg = shearing stress at the wall a, Further examples 217 stated that in regions of retarded potential flow the approximate solution becomes somewhat inacenrate as the point of separation is approached. The position of the point of separation can only be caloulated with a certain degree of uncertainty, parti ularly in cases when the point of separation is situated compnratively far behind the point of minimum. pressureft. From the assumption that the velocity profiles constitute a one-paruneter family it necessarily follows that the point of separation is determined solely by the value of this parameter. It was, however, shown by L, Tani [22] that. the position of the point. of separation depends, in addition, on the pressure gradiont of the external flow, d. Further examples In this section we propose to summarize some examples illustrating the calcula- tion of houndary layers by the preceding approximate methods which were first given ina paper by IL. Schlichting and A, Ulrich [19]. The first set of exaimples is concerned with elliptical cylinders whose major axes are parallel to the direction of the stream, ‘The ratio of the major to the minor axis of the eylinders ranged over a/b = 1,2, 4, 8 and the potential velocity-distribution functions arc seen plotted in Figs. 10.9. The value of the velocity maximum is U,,/U = 1 + ba. ‘the charac- teristic parameter of the boundary layer, namely the displacement thickness, 51, the shape factor, A, and the shearing stress at the wall, to, are seen plotted in Vig, 10.10, The results for tlie flat plate at zero incidence have been plotted in the same fignre for the purpose of comparison. In the case of a circular eylinder separa- lion oceurs at a/U’ = 0-609, i. ¢. at. 6 = 109-5°, as already mentioned, (2U = cirenm- ference) and moves downstrean: as the ellipse becomes more slender, The position of the point of separation is marked in the velocity profile plots in Vig, 10.9, The results for an ellipse of a/b = 8 differ only very little from those for a fiat plate at zero incidence. Fig. 10.11 contains velocity profiles for the boundary layer on an clliptic eylinder with a/b == 4. Calculations conceruing elliptic cylinders whose minor axes are parallel to the direction of the stream as well as ellipsoids of re- volution may be found in a paper by J. Pretsch [17]. : t Vor example: @, It, Schobaner [21[ measured the position of the point of minima prossare on an elliptical cylinder of slondemess @: 5 -= 290; 1 placed in a stream parallel to the major axis, Ho found that it was located at 2/6 = 1-3 and that separation took place at 2/6 = 1-9 A calevlation based on Pohthapsen’s approximation showed very good agresment with measvrements for velocity profiles np to the point of minimum pressure hut it predicted no jo». 1D, Meksyn [13} developed a method of compntation which leads to a valve of + 2-02 for the point of separation in the above example, by his method, the boundary layer equations aro transformed into ordinary differential equations which are related to Falkner and Skan's oqpration (9.8). t Here it may be worth mentioning that approximate integration hy the method of isaclines in connexion with Poblhausen’s approximation fails in the rogion of large presspre gradionts whiel occur for A > 12 (K > 0-095), becavse the plot of K againat A tims at Uda point (Table 10.2) and camot, therefore, be continucd heyond K - (095, Moreover, for A > 12 the velocity profiles become unacceptable as they contain points for which u/U7 ~'1 (Fig. 10.4). ‘These difficulties are obviated when eqn. (10.37) is sec. 218 X. Approximate methods for stendy equations A further example is shown in Fig. 10.12 which contains results for a symmetrical Zhukovekii acrofoil at: zero incidence. The point of minimum pressure is at x/i” = = 0-141 which is very far forward on the aerofoil. The pressure rise at the rear is very gradual so that the point of separation lies very far downstream of the point of minimum pressure, i.e. 2/l’ = 0-470. Since the Zhukovskii aerofoil has a ensped trailing edge the potential velocity at the trailing edge is different from zero. For details of additional systematic boundary-layer calculations concerning an extensive series of Zhukovskii acrofoils with different thickness and camber ratios and at different angles of incidence, reference may be made to a paper by K. Bussmann and A. Ulrich [2). Potential velocity distribution function on cllip- tical cylinders of slenderness ajb = 1,2,4,8, the direction of the atream being parallel to (he major axis So porition of point of separation ocular cylinder : ye r 06 08 x 10 ! x C Uv Fig. 10.10, Results of the calculation of houndary layers on elliptical cylinders of slenderness a/b = 1,2, 4,8, Fig. 10.9.0) displacement. thickness of the boundary layer, b) shape factor o) shearing stress al the wall, 21’ = circumforence of the ellipse; a/b — Lcireular oylinder; a/b = flat plate 4. Further examples 219 A review of the very numerous approximate methods which have been pro- posed so far is contained in the collective book entitlod “Laminar Boundary Layers” [18] and edited by 1. Rosenhead, In an effort to improve the accuracy of the calculation of laminar boundary layers, many authors replaced the preceding single-parameter methods by one em- ploying two parameters. This is achieved when the energy integral equation is satis- fied in addition to the momentum integral equation (sce e.g. K. Wieghardt: [27]). Two-parameter methods have been extensively developed by 1. G. Loitsianskii and his coworkers [8, 9, 10, 11, 12, 14]. Since such two-parameter methods are very complex, aud since their accuracy is difficult. to assoss, modern authors favour eaact numerical methods employed in conjunction with large electronic computers; their principles have been outlined in See. IXi. Fig. 10.11. Velocity profiles in the laminar Fig. 10.12. Velocity profiles in the laminar boundary layer on an elliptical cylinder. boundary layer aud potential velocity fune- Ratio of axes a/b = + tion for a Zlukovskii aerofoil 1.015 of thick- nest ratio d/l = 0-15 at an angle of incidence «=0 220 X. Approximate methods for steady equations e. Laminar flow with adverse pressure gradient ; separation Flows with adverse pressure gradien(s (retarded flows) are of great practical importance. In thin connexion itis alway donired to nid reparation trom the wall because this phenomenon is associated with large energy lorses. ‘The flow about an acrofoil is a casc in point. Owing to the fact that on the suction side the pressure must increase to its froe-stream valuc at the trailing edge, the flow is always likely (o separate. ‘The flow in a divergent channel (diffuser) affords another cxample. Tho object in using (his shape of channel is (0 convert. kinclie energy into pressure cnorgy, and if the angle of divergence is made too large, separation may occur. ‘Theoretien! investigations on the behaviour of the boundary layer in the vicinity of the point of separation have been carried out. by S. Goldstein [4] and 1.'S. Stratford [21a]. Cf. the review by 8, N. Brown and K, Stewarson |1 |, Observations show that a laminar boundary layer which separates from a wall frequently becomes reattached to it, having first become turbulent. This leads to the creation of a laminar separation bubble, Fig. 10.13b, which places itself betarcen (he separation point S and the reatt- achment. point. 2. ‘The fluid in (he bubble performs a circulatory motion. According to Fig. 10,13a, the pressure distribution along the wall can be represented, in simplified fashion, by a constant value between the point of separation S and point. 7' of largest thickness followed by a linear increase from T to the point of reattachment R. Phenomena of thie kind have been do- scribed in detail by 1. ‘Tani [23]. More recent experimental investigations into the nature of laminar separation bubbles have been performed by A. 1, Young et al, {28} as well as by M. Gaster |4a} and J. L. Van Ingen [6}. For theoretical contributions sce [2b, 3a, 5a). 14 will now be shown wilt the aid of several examples that a laminar flow ean only support very small adverse pressure gradien(s withont separation, Adverse presanre gradients which exist in practical applications would, therefore, almost always Ica (0 separation if the flow were laininar. The cirenmstance that real flows cau support con ble rates of pressure increase in a large number of cases without scparation is duc to the fact ¢hat the flow is mostly turbulent. It will be scen later that Gurbulent flows are capable of overcoming much larger adverse preaaure gradients withont separation. ‘The best known examples include the cases of flow past circular cylinders and spheres, when separation ocenrs inuch farther upsiream in laminar than iu tur- bulent flow. Jn practice when adverse prossure graclicnis exist. the flow is almost always turbulent hecanse, in addition, the cxistenee of an adverse pressure gradient favours the transition from laminar (o turbulent flow. It is, nevertheless, useful to clarify some of the fundamental relations assoriaed with the prevention of separation on the example of laminar flow, in particular, because lamin flows are much more readily amonable to thematical treatment than is the case with Curbulent flows. ‘There are several methods of preventing separation, The simplost of them consiste in anging for (he adverse pressure gradients to remain below the limit for which separation Fig, 10.13. Separation bubble in a laminar bondary layer after I. Tani 23}. a) Shape of bubble (nchematie) ; b) Pressure distribution in bubble along the wall (sche- matic), The préssure between S and V in (he bubble romaine constant at 74; further downstream the pressure increases to Pr S — point of separation R — point of reattachment Po T= night of turpbie c. Laminar flow with adverse pressur gradient; separation 221 does occur, A numerical example will serve to make this idea clear. Another possibility consists in rontrolling the boundary layer, c. g. by suction or by injecting Muid into it, or by addition of .n ncrofoil at a point where its presence favourably affects the boundary layer in critical regions, These methods will be discussod more fully in Chap. XLV, Following L. Prandtl (16) we shall show how it is possible to estimate the permissible magnitude of the adverse pressure gradient for which separation is just preventod. ‘The argument, will be based on the vou K&rmén-Pohlhausen approximation disenssed in See, Xb, HUwill beasoumed that the bowmdary layer is acted upou by the pressure distriby rorined by tho fe potential flow up toa point which lies ‘very close to the point, of separation, sich as point 0 ia ig. 10.14, Starting with this point it will be assumed that the pressure gradient is such that the © of the velocity profile remains unchanged proceeding downstream, or that, in other words, the shape factor A remains constant: since al separation = —12 a value of A 10 will be chosen. As seen from ‘Table 10.2 this leads to a definite value for the second shape factor, namely A = ~ 0-1369, so that P(X) = 1:523, Using trese values it is scen from eqns. (10.28) and (10,29) that. prevention of separation implics the following relationship between the velocity U (2) of potential flow and the momentum thickness 4,(z): 3? ow Tt followa that dZ/dz = 0-1369 U"/U"2, or a uu" 3 U Gy = 01809 Gag = 013690, (10.38) whore (10.39) UG), Fig. 10.14. Development of boundary Fig. 10.15. Potential velocity function layer in the case when laminar separation for 4 laminar boundary layer with aud is prevented without separation On the other hand the succeeding velocity profiles are given by the momentuin equation (10.36) for a > 0, or (10.40) where the nunerical valuo for F(K) which corresponds to A = — 10 has been inserted, From equa, (10.38) and (10.40) it follows that the value of the shape factor remains constaut at A = ~ 10 if 0:1369 7 = 1-523, or if uw “ye @ > Ino soparations 7 < IL: separation . (0.419) = IB =I, (10.41) The preceding argument shows that the boundary layer ean support the adverse pressure gra- dients if 6 > IL, whereas @ < LL implics separation. Jf ¢ remains constant at o = 11, sith A ~ ~ 10, the boundary layer reimains on the verge of separation, 222 X. Approximate methods for stendy equations Qualitatively it is at once possible to make the following statement regarding the shape of the potential velocity funtion U(x) which leads to no separation, In viow of eqn. (10.41) ov>o is a necessary condition for a retarded flow (U/" < 0) to adhere to the wall. In other words, the magnitude of the adverse pressure graciont must decrease in the flow direction. Mig. 10.1%. ‘Thus separation will always occur il the function (/(z) is curved downwards behind ite maxinnin (U" 0), separation may be obviated, ven the limiting case of 7” = 0,i.e. the case of a velovity which decreases linearly with the length of are, always leads to separation, This latter remark agrees with the result found in See. IXd; it was concerned with the boundary layer associated with a potential flow velocity which decreased linearly, and the solution of the dillerential equations was quoted from a paper by J. Howarth. The su/ficient condition for the absence of separation is given by Uv" > tbo. Wo shall now proceed to calculate the potential flow and tho variation of boundary-layer thicknoss which are associnted! with the limiting case of o == L1, when tho boundary layor rensains on the verge of separation, Jrons eqn. (10.41) we have wu uv Ux a or, upon integrating: In U’ = E1 In J -L In (—@,'), i, UU" = ~ CY, whore OC,’ denotes the constant of integration. Repeated integration’ gives ob U- = G2 + Cy. (10.46) For x = 0 wo shot! have U(2) == Ug, 80 that Cy = yyUy-™. Putting further C,’ Us" = Cy, we obtain from equ. (10.41) U(x) = (10.43) Equation (10.43) represents the potential velocity for which separation caw just be avoided. ant C, cas he determined from the value of the boundary-layer thickness dy at te . Wo have A = U’ 5%/v = — 10 or 5 = 10 ¥/(—U"). From eqn. (10.43) we obtain ” C, Uy = — fod, xii and henee . io 11720 Gy Et HOC From 6 = dy at x = 0 we have 0, = 10 1/Up 59%, which gives the final solution for the potential flow and the va intion of boundary-layer thicknoss U(r) = Uy (1 +100 (10.44) we \O58 ste) = 4 (14 1005°% 2) . (10.45) Jt is sen that. the magnitude of the permissible deceleration (decrease in velocity) is very small, being proportional to x 1, Its value is very nearly realized for the enso of constant velocity . In the present case the inerease in boundary-layer thicknoss, this vnine also differs but little from the ease of a flat plate at zero 5. proporticn lenee for whieh 5 ~ a! References 223 By way of a further example of rotarded flow we shall consider the flow through a divergent channel whose walls are straight. ‘This case is corollary to the caso of the houndary lnyor in a divergent channel trent in Seo. IX b. ‘The flow is acen sketched in Fig, 10.10, whore x denotes the radial distance froin the source at 0, The wall is assumed to hegin at x = a where the entrance velocity of the potential stream is put equal to Up. The potential flow is given by Ug =U Ss Vix Ss Ue) =2, 5. (10.46) Computing the value of the quantity ¢ from eqn. (10.41), which is decisive for soparation, we obtain here ¢ Applying the criterion given in eqn. (10.412) we conclude that separation ocenrs in all cases irrespsctive of the magnitude of the angle of divergeuce. ‘This cxample slows very clearly that ¢ laminar stream has only a very limited capacity for supporting an adverse pressure gradient without separation. According to a calculation performed by K. Pohlhansen {15} the point of separation occurs at x,/a = 1-21 and is seen to be independent of tho angle of divergence. Fig. 10.16. Laniinar boundary layer in a divergent, channel, Separation occurs at x4/a = L-2E indepen. dontly of the angle of divergence ‘The preceding conclusions apply only as long, os the displacement effect of the boundary layer may be neglected. However, this is not the case when the angle of divergence is sinali, ‘When this angle is small, the boundary layers fill the whole channel cross-section after a certain inlet longth (c). See. XL i) and the flow goos over asymptotically to that discussed in Sec. V 12 under the hearing of channel flow. When the included angle docs not exceed a certain value which depends on the Reynolds number, there is no separation. Recently. 8. N, Brown and K, Stewartson {1] published a synmary review om separation in which the mathematical question centered on the singularity which oeeure in the dill equations at the eritien! point has been enrphasized, Sce also the work of 8. Goldstein [4]. A more physically inspired review of thin problem area has recently been published by J.C. Williams M11 ]29}, and by P. K. Chang {2c}. References {1} Brown, §.N,. nnd Stowartson, K.: Laminar separation. Annual Review of Phaid Me 45 —72 (1969). 12] Bussmann, K., and Ulrich, A.: Systenntisehe Untersucinngen fiber den Kinttuss der Profil: form auf dic Lage des Uinschlagpmktes, Preprint dh. at. Loftfahrtforscliaig 103 in: ‘Techn. Rerichte 70, No. 9 (1942); NACA TM 1185 (1947), J2a] Chan, Y.Y.: Loilsiauskii's method for honudary layers with suction and inje 7. 502-5 (1060), {2b} Briley, W. 4%, and Meponald, D. 7 bubbles. JEM 69, 631—656 (1975). [2c] Chang, P. K,: Separation of flow, Pergamon Press, New York, 1970. [3] Glanert, M.B., and Lighthill, M.J.: The nxisyimnetric boundary layer om a long thin cylinder, Proc. Roy. Soe. A 230, 188-203 (1955). [3n]} Crimi, P.. and Reeves, B.L.: Analysis of lending edge soparation bubbles on airfoils. AIAN J. 14, 1548 ~ 1555 (1976). [4] Goldstein, 8.: On Imninar boundary Inyer flow ne: Appl. Math, 7, 43-60 (1948). 1 on. ALA Numerical prediction of incompressible separation a point of sepuration, Quart. J. Mech. 224 X. Approximate methods for steady: equations [da] Guster, fhe structure and behavionr of laninny separation bubbles, AGARD Conf. Proce. 4, 819- 854 (1961 }5] Holstein, 11, and Bol achielten, die dem Ni S. 10, 5 "16 (1940) [sn] Horton, 11, Pz A semi-empirical theory for the growth and bursting of laminar separation bubbles. Acro. Res. Council, Carrent Paper No. 107 (1967). [8] Ingen, -IeL.t On the ralenlation of lami ration bubbles in Qvedimens incompressible flow. ACARD Coul, Pro Wa TE 1 to Tt (1975). 17] vou Kérmin, ‘The: Oher lauinare nl Grebatente Reibung, ZAMM 7, 233 262 (1921 . Works 11, 70-97 (1956). ty Neal. and Lait 1. G.: Ober eine angeniilierte Method der Rereehuung, uminargrenzschicht. Dokl. Akad. Nauk, SSSR 36, No. 9 (1942); see nlso: Ans approxi- wate method of calewlating (lie Im : Kin einfaches Verfaliren zur Berechumug laminarcr Reilmugs- rngsverfaliren von K. Polilhausen yemigen, Lilienthal. Bericht Lautinnrny? pogranichnyi sloi, re Grenzschichien, Akademie: Verlag, Berlin, 1967, Guz Mekhanika zhidkostei i gazov, Nauka, Moscow, 1973. Universal'nye uravnenia i parametricheskie pribl nykh pogranichnyh slocv, Prikl, Mat, + Mekh, XX/X. No, 1 (1965 oquations and parnmetric npproximati Math, Mech, (PMM) 29, 70-87 (1965). [12| Loitsinns Sur In méthode pnrandétrique cle Proc, 11th Intern, Congress Appl. Moch., Mimieh 1 Berlin, 1900, 722-728. (13) Nokayn, 1.3 Integration of (te boundary hiyer equations. Prov, Roy, Sor, A 237, 643 059 956), [14] Ozerova, 15... and Si for stondy-state laminar boundary fa (£970). [15] Pohlhausen, K.: Zur nihernngeweisen Integration der Diflerentialgleichnng der laminaren Reibmngsechicht. ZAMM 1, 252-268 (1021). [16] Prandtl, L.: ‘The mechanics of viscous flnids. In W.P, Durand (ol): Aerodynamic Theory TI, 34~ 208 (1939). {17] Pretach, J.: Die laminare Reibungsschicht an clliptischon Zylindern und Rotationsellipso- iden bei syminetrischer Anstrommng, Luftfahrtforschung 18, 307—402 (1941). [18] Rosenhead, 1. (ed.): Laminar boundary layers. Clarendon Press, Oxford, 1963. [19] Schlichting, H.. and Ulrich, Ac: Zur Berochnung «des Umschlages laminar-tirbulent. db, dt. Larftfahriforschnng 1, 8--35 (1942); see also: Litionthal-Berieht $10, 75-~135 (1940). [20] Schénan i Differenzenverfahren zur Lisung der Grenzschichtgleichnng fitr statio- hare, Inuinare, inkomprossible Stramung. Ing.-Arch. 33, 173~ 189 (1964). B.: Airflow in a separating laminar boundary layer. NACA Rep. 527 (1935). Flow in the laminar boundary Inyer near separation, ARC, RM 3002, 1-27 19} Loitsian HL. Liber lintite huninaire, Springer Verlag, nimi, L.M.t Approximate (0 yers (in Russian). 1 rameter solution of the equal it Polyt. Inst. No. 313 {22} Tani, 1 fly years of bomvlary On the solution of the laminar bonndary layer equations. layer research (I1. Girtler, ed), Brannachweig, U [23] Tani. 1. Low specd flows involving bubble separation. Progress in Acronnutical Sciences 5, 70-103 (1964), [24] Truckonbrodlt, Eo: Nihermgslésungen der StrOmungsmechanik und ihre phy: Dentung. Nineteonth Prandti Memorial Lecture, ZFW 24, 177—188 (1976), [25] Walz, Av: Bin noner Ansatz. fiir das Gesclwindigkeltsprofil der laminaron Reibungssehicht, Lilienthe nt 141, 8-12 (1941), 1 [26] Watson, W.1,, and Preston, J. FL: Au approximate solution of two flat pla layer problems, ARC RM 2537 (1951). 127] Wieghardt, K.: Ober cinen Energicsa tz zur Berechwung laminarer Grenzachichten. Ing.- Arch. 16, 231 242 (1948). [28] Young. A.D., and Horton, 1.P.; Some results of investigations of separation bubbles. AGARD Conf, Proc. Mow Separation 4, Part 1, 779--818 (1966), [29] Williams, M1 Tucompressible boundary layer separation, A Mech, 1113 144 (1977). lische © boundary ew of Fluid nin Rev CHAPTER XI Axially symmetrical and three-dimensional boundary layers. In the discussion of boundary layers in the preceding chapter we have considered exclusively two-dimensional cases for which the velocity components depended on only two space coordinates. At the same time the velocity component in the direction of the third space coordinate did not exist. The general three-dimensional case of a boundary layer in which the three velocity components depend on all three co- ordinates has, so far, been hardly claborated because of the cnormous mathematical difficultics associated with the problem. We shall describe the first attempts in this dircetion at the end of the present chapte On the other hand the mathematical diffienlties encountered in the study of axially symmetrical boundary layers are considerably fewer and hardly exeeed those in the two-dimensional case, Axially symmetrical boundary layers oenr, ¢. g.. in flows past axially symmetrical borlies; the axially symmetrical jet also belongs under this heading. ‘Iwo cxainples, that of the rotating disk and axially symmetrical flow with stagnation, have already heen disenssed in the chapter on oxact solutions of the Navier-Slokes equations. We shall begin the present chapter with a discussion of some further examples of steady axially symmetrical flows which can be solved with the aid of the differential equations, and will continue with the extension of the approximate procedure which was explained in the preceding chapter to include the axially symmetrical ease. Furthor, we shall discuss the principal features of three-climensional boundary layers, Non-stcady axially symmetrical boundary layers will be considerod in Chap. XV together with non-steady two-dimensional examples. a, Exact solutions for axially symmetrical boundary layers 1. Rotation near the ground. In Chap, V we have considered the ease of flow in the neighbourhood of a disk which rotates in a fluid at rest. ‘The case of motion near a stationary wall, when the fluid at a large distance above it rotates at a constant angular velocity, is closely connected with it, Fig. L.1. This example was studied by U. T. Bocdewadt [9]. One of the essential effects in the example of the disk whiclh rotates in a fluid at rest consists in the fact that in the thin layer near the wall the fluid is thrown outwards owing to the existence of centrifugal forces, The fluid which is forced outwards in a radial dircetion is replaced by a fluid stream in the axial direction, In the case under consideration, in which the fluid rotates over the wall, there is a similar effect, but its sign is reversed: the particles which rotate at a large 226 XL. Axially aymmoteical and three-dimensional boundary layers distance from the wall are in equilibrium under the influence of the centrifugal force which is balanced by a radial pressure gradient. The peripheral velocity of the particles near the wail is reduced, thus decreasing materially the centrifugal force, whereas the radia! pressure graclient directed towards the axis remains the same. This set of circumstances causes the particles near the wall to flow radially inwards, and for reasons of continuity that motion must be compensated by an axial flow upwards, as shown in Fig, 11.1. A superiinposed ficld of flow of this nature which occurs in the boundary layer and whose direction deviates from that in the external flow is quite generally referred to as a secondary flow. It was first dis- covered by FE. Gruschwitz [45] when he analyzed the flow in a curved channel, sce also E. Becker [6]. Fig. LL.1, Rotation of flow near the ground Velocity componente: — rade; e—iangentiat; w—axlel, Owing to friction, the tangential vetoclty sorters a deceleration In the nelghbourhood of the disk at rest, This gives rise to a necon- dary flow which In Aivected cedtally tn- wards The secondary flow which accompanies rotation near a solid wall and which hes been described in the preceding paragraph can be clearly observed in a teacup: after the rotation has been generated by vigorous stirring and again after the flow has been left. to itself for a short, while, the radial inward flow field near the bottom will be formed. Its existence can be inferred from the fact that tea leaves settle ina little heap near the centre at the bottom, In order to formulate the mathematical problem, we shall assume cylindrical polar coordinates r, $, 2, the stationary wall being at z , see Fig. 11.1. The fluid at a large distance from the wall will be assumed to rotate like a rigid body, with a constant angular velocity @. We shall denote the velocity componcnt in the radial direetion by u, that in the tangential direction by v, the axial component being denoted by 1. For reasons of axial symmetry the derivatives with respect to ¢ may be dropped from the Navier-Stokes equations. The solution which we are about to find will be an exact solution of the Navier-Stokes equations, just as was that for the a, Exact solutions for axially symmetrical boundary layers 227 rotating disk, because the terms which are neglected in the boundary-layer equations vanish here on their own accord. By eqn. (3.36) we can wrile down the Navier-Stokes equations as ue bw He (I1.Ja) uw Ne (11.1b) ow a ue bw ie (IL...) (111d) The boundary conditions are z=0: u=0; v=0 ; w=0 | (11.2) za00: u=0; v=re. It is convenient to introduce the dimensionless coordinate part (11.3) in place of z, as in the ease of the rotating disk (See. V 11), We assume that the velocity components have the form u=roP(); v=reG(); w=V~roH(. (11.4) The radial pressure gradient can be computed for the frictionless flow at a large distance from the wall from the condition: (1/e) - (@p/ar) = V2Jr, or, with V = ra, Lap _ : oe ara. (11.5) e In the framework of the boundary-layer theory it is assumed that the same pressure gradient acts in the viscous layor near the wall. Introducing eqns. (11.4) and (11.5) into eqns. (11.14, b, d), we obtain a system of ordinary differential equations which is analogous to that in See. V 11: F?-_@? 4. HF’ — F"+1=0 26F+HG =0 (11.6) 2F 4-H’ =0 with the boundary conditions ¢=0: F=0; G=0; H=0; | (11.7) C00: F=0; G=l1. 228 XI. Axially symmetrical and (heee-dimensional boundary layers The pressure gradient in the z-direetion may be assumed equal to zero, as such an assumption is compatible with boundary-layer theory. Alternatively, it can be calculated from eqn. (Hi.f¢), after the principal solution had been obtained, whiclr then results in an exact solution of the Navier-Stokes e ions. The system of equations (11.6) with the boundary conditions (11.7) was first solved by U. T, Bocdowadt [9] ina very lnborious way by meuns of # power series ox- pansion at 2 = 0 and an asymptotic expansion for ¢ =: co. Recently this sultion was corrected by J.B. Nydahl[81a} in an unpublished paper, The values of the functions F,G, TT according to Nydahl are given in Table 111 and in Fig. 11.2. The horizontal velocity. i.e. the resultant of w and 9, is also shown plotted ina polar diagram in ig. 11.3. Tho angle between the horizontal velocity component and the peripheral direction depends only on the height, and the vectors in Mig. 11.3 indicate this direction for varying icights. ‘The deviation from the peripheral direction prescribed ata large height is largest near the ground and has a value of 50-6? inwards. ‘The largest. deviation of 7-4° ontwards occurs for ¢ = 463 so that the largest angular Table 11.1, The functions for the velocity distribution for the case of rotation over a . stationary wall, after J, E, Nydahl [8la} é G H 0.0 0.0000 05 0.3487 1.0 0.4788, 15 0.4496 0 2.0 —0,3287 . 2.5 0.1762, 1.7459 3.0 0.0361 1.8496 3.5 0.0663. 1.8308 4.0 0.1227 1.7325, 45 0.1371 5.0 0.1210 5.5 0.0878 6.0 0.0499 65 0.0162 7.0 = 0.0084 15 0.0223 8.0 0.0268 85 0.0243 9.0 179 9.5 1.0102, 10.0 0.0033, 10121 1.3683 10.0 0.0018 1.0099 1.3689 11.0 0.0047 1,0065, 1.3654 11.5 0.0057 1.00381 1.3601 12.0 0.0052 1.0003 12.5 0.0038 0.9984 7 09,0000 1.0000, a, Exact solutions for axially symnictrical boundary layers 229 20 18 1S 6 uw |-—]—-| -\Y- ||| eo 12 | || | » fr Pe, FY cicumerenat 7%, ection "2 °m, 08 . aiecton pn urediat Jowards axis Fig. 11,3. Rotation near a solid wall, after Boedewadt, Vector representation of the horizontal velocity comportent ~< Fig. 11.2. Rotation near a solid walh after Boedewadt [9]. Velocity distribu. tion in the boundary layer from egn. (11-4); nce alao Table 11.1 difference, i. ¢. that between the ground and that at ¢ = 4-63, is 58°. It is further remarkable that the axial velocity component w does not depend on the distance r from the axis but only on the distanee from the ground. The motion at all points upwards with w > 0. As already mentioned, this is caused by the inward flow near the ground, consequent upon the deercase in the centrifugal forees. In any ease, as scen from Fig, 11.2, this is compensated hy a radial flow ontwards at a greater height, but on the whole, the radial flow inwards predoriinates. The total volume towing towards the axis taken over a cylinder of radius R around the 2-axis is Q=2nR jude = 2nR? Yor [Fe dt =: —2R? fw (co). é 250 Inserting the nimerieal value of [1 (eo) from ‘Table 11,1 we obtain Q = — 13872 Vw. (1.8) ‘The volume of flow in the positive z-direction is of equal magnitude. The largest upward motion occurs at € == 3-1, where w == 1-85 /@ 7. Tt is also worth noting that. 230 XI. Axially symmetrical and three-dimensional boundary layers the boundary layer extends considerably higher than in the example with the disk rotating in a fluid at rest (Sec. Vb). If the boundary-layer thickness 6 is defined as the height for which the deviation of the periplicral velocity is equal to 2 per cent, we shall obtain 5 — 8 ¥ v/a as against 5 = 4 Yo] for the stationary fluid. The cxample of the motion of a vortex source between two parallel walls con- sidered by G. Vogelpohl [120] is related to some extent to the present ease, For very small Reynolds numbers the velocity distribution deviates little from the parabolic curve of Poiseuille flow. For Inrge Reynolds numbers the velocity profile approaches a rectangular distribution, and a boundary layer is seen to be forming. The corresponding case of turhulent flow was discussed by C. Pfleiderer [85]. In this connexion the paper by E. Becker [6] may also be consulted. Similar phenomena can be found in swirling Now through a conical funnel-like channel investigated by K, Garbsch [32]. The potential flow ie generated by a sink of strength Q placed at the vertex of the cone and a potential vortex of strength 1° placed along the axis, Fig. 11.4, The solution to the boundary-layer equations is obtained by an iterative procedure which is baid to lead to a good approximation with a small number of slope only. ‘Two particular enens of such flown have also heon investigated with the aid of approximate methods, and they will be mentioned in Chap. X: A.M. Binnie and D, P, Harris (7] studied pure sink flow (1” = 0), and G, I, Taylor [111] and J.C, Cooke [17] stndicd pure vortex flow (Q = 0). In the latter case, aa shown in Hig. 11.4, the flow forms a boundary lnyer on the wall of the conical channel, The flow field in tho boundary layer develops a velocity component. w in the direction of tite cone generators, whereas tte frictionless corc, being a pure swirl, possesses only tangential velocity components v. ‘The sccondary flow in tho boundary layer transports some fluid towards the vertex. The reader may further wish (o svady a related paper by H. E. Weber [121]. 2. The circular jet. We shall now itvlicate 11, Schlichting’s [97] solution for the laminar circular jot which is analogous to the one for a two-dimensional jet given in Sec. 1Xg. The subject of the investigation is, thus, a jet which leaves a small circular opening and mixes with the surrounding uid, In most practical cases the circular jet is also turbulent, ‘The turbulent circular jet will be considered in Chap, XX1V, but since it leads to a differential equation which is identical with that for the laminar case we shall discuss tho latter in some greater detail. Vig. 11.4. Swirling flow in a convergent conical channel, af. tor G.I. Taylor [111] I = boundaey layer on the wall o¢ the conteal channel with secondary flow towards the vertex a, Exact solutions for axially symmetrical boundary layers 231 The pressure can here be regarded constant, as in the two-dimensional case. The system of coordinates will be selected with its x-axis in the axis of the jet, the radial distance being denoted by y. The axial and radial velocity components will be denoted by u and », respectively. Owing to the assumption of a constant pressure the flux of momentum in the direction of x is constant onee more: J = 2g [uy dy = const. (11.9) é In the adopted system of coordinates the equation of motion in the direction of x, under the usual boundary-layer simplifications, together with the equation of motion, can he written as ou ou SE oe (1L.10a) ou Ed (11.10b) and the boundary conditions are =0: v=0; y | (iL) yoo: u=0. As before, the velocity profiles u(x, 4) can be assumed similar, The width of the jet will be taken to be proportional to 2, it being further assumed that y a pwr Fn) with = In order to determine the exponents p and m we can use the same two conditions as in the two-dimensional case. First the momentum from eqn. (11.9) must be indepen- dent of x, and secondly, the inertia and frictional terms in eqn. (11.104) must be of the stine order of inagnitude. Hence au au la a pe ern in, SE ptm ptm Ey SE) gan, um arrin, ye (y ) a Thus the following two equations for p and n result: Qp—4n4f2n=0; B2p—4n— so that p =n = 1. Consequently, we may now put yp=r2F(y) and y= g : from which it follows that the velocity components are Fy umd (r- *) (1.12) us ale 232 XI. Axially symmetrical and throe-dimensional boundary layers Inserting these values into eqn. (11-102), we obtain the following equation for the stream function FEOF? PFT od ( " ") CRS PRO (er FP), ue ” " 1 7 which can he integrated onec to give FF =F’ —7F". (1.13) The boundary conditions are x =m, and v = 0 for y = 0. It follows that F” =: 0 and F =0 for 9 = 0. Since 1 is an even function of 7, F']y must be even, F’ odd and F cven. Because of F (0) = 0 the constant term in the expansion of F in powers of 1 must vanish, which «letermines one constant of integration. The second constant of inte- gration, which will be denoted by y, can be evaluated as follows: If F (1) is a solution of eqn. (11.13), thon F(y 9) = F(é) is also a solution. A particular solution of the differenti! equation dF dF ar Pag ag ae which satisfies the boundary condition € = 0: F = 0, F’ = 0, is given by (iii4) More & = y y/x, and the constant of integration y can now be determined from the given value of momentum. From eqn. (11.9) we obtain for the momentum of the jet Jaane f ty dy = Buoys. Finally, the above resvlts can be expressed in a forin to coutain only the kinematic viscosity, 7, and the kinematic momentum K' = J/o. Thus 30K’ 1 wae dK od, (11.15) Bnvx (1 + i #) ' 1 1 yi fe v= aK * o (11.16) Va = Gate vy, (11.17) a. Bxact solutions for axially symmetrical houndary layors 233 Figure 11.5 represents a streamline pattern calculated from the preeeding equations. The longitudinal velocity x is shown plotted together with that forsthe two-dimen- sional jet in Fig. 9.13. ‘The volume of flow Q=22 fy dy (volume per second), which increases é with the distanee from the orifice owing to the flow from the surroundings, is repre- sented by the simple equation Q=8xve. (11.18) Fig. 11.5. Streamline pattern for a circular laminar jot ‘This equation should be compared with eqn. (9.48) for the two-dimensional jet. It is seon that, unexpectedly, the volume of flow at a given distance from the orifice is dlopendent of the momentum of the jet, i. c., independent of the excess of pressure under which the jet lenves the orifice. A jet which leaves under a large pressure difference (Inge velocity) remains narrower than one leaving with a sinaller pressure difference (small velocity). The latter carries with it comparatively more stationary fluid, namely in a manner to make the volume of flow at a given distance from the orifice equal to that in a faster jet, provided that the kinematic viscosity is the same hoth cases, IL. B. Squire [105, 106] was able to find solutions to the boundary-layer equations as well as to the complete Navier-Stokes equations and to inake u comparison bebween them for the case of a conical jot. v possesses an additional, radiul velacit ponent in its annular oi Tn this latter class ef radial jets the velo inversely proportional to the distance froin the orifice. ‘The theory can be extended to turbulent flows by replacing the kinematic viscosity with the apparent kinematic viscosity of turbulent flow, which in this ease remains constant, see Chap. XXIV. The case when a jet impinges at right angles on a wall and is spread along it, was solved hy M. B. Glanert [40], who included plane as well as axially symnictrical, and luninar as well as turbulent flows. com. The corresponding ease of a compressible circular laminar jet was evaluated by M. Z. Kraywoblocki [61] and D.C. Pack [83]. In the subsonic regime, the density on the axis of the jet is larger, and the temperature is smaller than on its boundary. These differences are inversely proportional to the square of the distanee front the 234 XL. Axially aymmetrical and three-dimensional boundary Inyers orifice. According to TH. Goertler [43], the case when a weak swirl is superimposed on the jet can also be treated mathematically, and the effect of the swirling motion present in the orifice ean be. traced in the downstream direction. Tt turns out that the swirl decreases faster with the distance from the orifice than the jet velocity on the axis. 3, The axially symmetric wake. ‘The flow in an axially symmetric wake, such as occurs downstream of an axially symmetric body placed in a stream parallel to its axis, can also be described with the aid of the aystem of equations (11.10, b). The solution is quite analogous to that. for the two-dimensional case which was de- scribed in detail in See. IXf, Let Uso denote the oncoming velocity and let u(x, y) be the flow velocity in the wake, We assume, as was donc in eqn. (9.29), that the velocity difference in the wake, U, (x,y) = Uco ~- u(a,y) (11.19) is very small compared with U., far downstream. Consequently, we shall neglect quadratic terms in w,. With this simplification it is possible to deduce from eqns. (11.108) and (11.19) the following differential equation for 1: Ou, v d Ou, Un Me 2 (y MH). (11.20) The analytic form to be assumed for the dependence of the velocity difference 4u,(t,y) on the axial coordinate, x, and on the radial coordinate, y, can be discovered from the condition that the drag evaluated from the momentum of the wake must become independent of « at large distances downstream of the body, ‘This leads to the relation D=279 Ua fu-y dy = const, (11.21) v0 which is satisfied by the form y= cu, 1. , (11.22) nahy joe : (11.23) This form is analogous to thet in cqn, (9.31) for the’ two-dimensional problem. Substituting eqns, (11.22) and (11.23) into eqn. (11.20), we obtain a differential equation for {(1). This is (if) +29 f + 4nf=0, (11.24) where and the boundary conditions are ‘ Af % f=0 at 4=0 and f=0 at »=00. It is easy to verify that the solution of equ. (11.24) has the form of an exponential, 1(n) = exp (— 9), (11.25) a, Exact solutions for axinlly symmetrical boundary layers 235 this form, too, being analogous to that in eqn. (9.34) for the two-dimensional ease. Ifence, the velocity difference turns out to be uy (ny) = £ Unexr(— } fet) The value of the constant C must be determined from the drag with the aid of eqn. (11.21); its value is C= 3 coRd, where ¢p denotes the drag coefficient referred to the frontal area of the body, and R= Udy, Hence we obtain , aty [ad ey) fp (¢ k) exp (— 7). (11.26) Ue 32 The plot of the velocity difference from eqn. (11.26) is the same as that in Fig. 9.10. Experimental data can he found in F. R. Hama’s work [45a]. 4, Boundary layer on a body of revolution. The flow of a viscous fluid past a body of revolution when the stream is parallel to its axis is of great practical impor- tance. The boundary-layer equations have heen adapted to this case by B, Boltze [10]. Assuming a curvilinear system of coordinates (Fig. 11.6), we denote by « the current length measured along a meridian from the stagnation point, y denoting the coordinate at right angles to the surface. The contour of the hody of revolution will be specified by the radii r (x) of the sections taken at right angles to the axis. We assume that there are no sharp corners so that d?r/d2? docs not assume extremely large values. The velocity components parallel and normal to the wall will be denoted by wand v, respectively, and the potential flow will be given by U(x). According to Boltze the paalacetare equations will then assume the form: oe au L ap get ’ay mopar? ay (11.27a,b) with the boundary conditions: y=0: w= 5 yoo: n= U(x). (11.28) Fig. 11.6. Boundary Inyer near a body of revolution. System of coordinates 236 XI. Axially symmetrical and three-dimcnsional boundary layers ‘The equation of motion in the x-direction ia seen to remain unchanged compared with two-dimensional flow. An order-of-magnitude estimate of terms in the equation of motion in the y-direction shows that the pressure gradient normal to the wall aplay ~ x2/r ~ 1. Consequently the pressure difference across the boundary layer is of the order of the boundary layer thickness 5, and it is again possible to assuine that the pressure gradiont of the potential stream, @p/az, is impressed on the bound- ary layer. We shall lnnit the considerations of this chaptor to the case of steady flow. In order to integrate eqns. (11.27a, b) for the axially symmetrical case it is once more possible to introduce a stream function (x,y) given by: ub ae) oy ray "dy? (11,298, b) ft vp- tan @y dar SO Eon Oe de ¥ This transforms eqn. (11.271) into dp ay (3 lar ay au ay By dz ay \az b 7 de®) ay =F az 1” ays (11.30) with the boundary conditions a y= y=0, xo =0; y=00: #012). We now proceed to give a brief account of the methods used to calculate the boundary layer on a body of revolution. A more detailed account can be found in an earlier edition of this book [101]. The numerical results for a sphere, however, will be discussed in more complete detail. ‘The boundary layer on a body of revolution of arbitrary shape can be determined by the same method as that used in Seo. LX c for the case of a cylinder of arbitrary cross-section (two-dimensional problem). The velocity of the potential flow, U(x), is expanded into a power series in x and the streain-function, y , is assumed to be represented by a similar series in a, with coeffi- cients depending on the wall distance y (Blasius scries). Following N. Frocssling [29] it is found that here also the coefficient-functions of y can be so arrangec| as to become independent of the parameters of any particular problem, In this manner the fune- tions can be calculated once and applied universally. ; The equation of continuity can also be satisficd by an’ alternative stream function j, such that, wal 2h, yt 2% r Oy 7 Ox ‘This form of the streain function was uscd by BE. Boltze when he calculated non-steady axially symmetrical boundary layers, as described in Sec. XVb2.

You might also like