Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Separation of Flow
Separation of Flow
Separation of Flow
Ebook1,462 pages12 hours

Separation of Flow

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Interdisciplinary and Advanced Topics in Science and Engineering, Volume 3: Separation of Flow presents the problem of the separation of fluid flow. This book provides information covering the fields of basic physical processes, analyses, and experiments concerning flow separation. Organized into 12 chapters, this volume begins with an overview of the flow separation on the body surface as discusses in various classical examples. This text then examines the analytical and experimental results of the laminar boundary layer of steady, two-dimensional flows in the subsonic speed range. Other chapters consider the study of flow separation on the two-dimensional body, flow separation on three-dimensional body shape and particularly on bodies of revolution. This book discusses as well the analytical solutions of the unsteady flow separation. The final chapter deals with the purpose of separation flow control to raise efficiency or to enhance the performance of vehicles and fluid machineries involving various engineering applications. This book is a valuable resource for engineers.
LanguageEnglish
Release dateJun 28, 2014
ISBN9781483181288
Separation of Flow

Related to Separation of Flow

Related ebooks

Technology & Engineering For You

View More

Related articles

Reviews for Separation of Flow

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Separation of Flow - Paul K. Chang

    SEPARATION OF FLOW

    PAUL K. CHANG

    Professor of Mechanical Engineering, The Catholic University of America, Washington, D.C., U.S.A.

    Table of Contents

    Cover image

    Title page

    Copyright

    Preface

    List of General Symbols

    Chapter 1: Introduction to the Problems of Flow Separation

    Publisher Summary

    Symbols

    Subscripts

    Superscripts

    1 Mechanics of Flow Separation

    2 Modern Development of Flow Separation Theory

    3 Mechanics of the Separated Flow

    4 Analysis of Flow Separation

    Chapter 2: Steady Separation of Incompressible Laminar Flow from Two-dimensional Surfaces

    Publisher Summary

    Symbols

    1 Examples of Two-dimensional Steady Laminar Flow Separation

    2 Analytical Methods for Prediction of Laminar Flow Separation

    3 Discussion

    Chapter 3: Steady Incompressible Laminar flow Separation on Bodies of Revolution and Other Three-dimensional Solid Configurations

    Publisher Summary

    1 Three-dimensional Laminar Flow Separation

    2 Discussion

    3 Laminar Flow Separation on a Body of Revolution and a Delta Wing

    Chapter 4: Incompressible Turbulent Flow Separation

    Publisher Summary

    Symbols

    1 Calculation of Momentum Thickness

    2 Criteria of External Incompressible Two-dimensional Turbulent Flow Separation

    3 Internal Flow Separation

    4 Axially Symmetric and Three-dimensional Steady Incompressible Flow Separation

    Chapter 5: Unsteady Incompressible Flow Separation

    Publisher Summary

    Symbols

    1 Boundary Layer Separation after Impulsive Start of Motion

    2 Axially Symmetric Unsteady Separation

    3 Boundary Layer Separation with Constant Acceleration

    4 Laminar Separation from a Moving Wall

    Chapter 6: Compressible Flow Separation

    Publisher Summary

    Symbols

    1 Determination of Compressible Laminar Flow Separation Point without Interaction

    2 Nature and Characteristics of Shock-induced Separated Flows and their Influences Up- and Downstream

    3 Pressure Rise Involving Shock Interaction

    4 Analysis of Shock-induced Separation

    Chapter 7: Characteristics of Separated Flows

    Publisher Summary

    Symbols

    1 Separated Flows Over and Behind Two-dimensional Surfaces and Axial Symmetric Bodies

    2 Separation Bubbles

    Chapter 8: Wake Flow

    Publisher Summary

    Symbols

    1 Wake Flow at Subsonic Speeds

    2 Wake Flow at Supersonic and Hypersonic Speeds

    Chapter 9: Leading-edge Flow Separation

    Publisher Summary

    Symbols

    1 Leading-edge Flow Separation at Subsonic Speeds

    2 Criterion of the Laminar Leading-edge Flow Separation

    3 Bursting of the Leading-edge Vortices

    4 Leading-edge Flow Separation at Supersonic Speeds

    5 Flow Separation on Thin Protruding Probes placed in Front of Blunt Bodies at Supersonic/Hypersonic Speeds

    Chapter 10: Base Pressure

    Publisher Summary

    Symbols

    Subscripts

    1 Base Pressure at Subsonic Speeds

    2 Base Pressure at Supersonic Speeds

    3 Experiment on Base Pressure Behind a Two-dimensional Surface

    4 Analysis of Base Pressure Behind a Two-dimensional Surface

    Chapter 11: Thermal Effects on Separation of Flow

    Publisher Summary

    Symbols

    1 Physics of Heat Transfer in Separation

    2 Heat Transfer Effect on Separation at Subsonic Speeds

    3 Position of Compressible Flow Separation

    5 Analysis of Heat Transfer Effects on Separated Flow

    5.2 Chapman’s theory of heat transfer in regions of separated flow

    5.2.1 SOLUTION OF MOMENTUM EQUATION

    5.2.2 SOLUTION OF ENERGY EQUATIONS

    5.2.3 VELOCITY, ENTHALPY, AND TEMPERATURE PROFILES IN PHYSICAL COORDINATES

    5.2.4 HEAT TRANSFER AND RECOVERY FACTOR AT WALL

    5.3 Carlson’s analysis of laminar separated flow at hypersonic speeds [93]

    Chapter 12: Control of Separation of Flow

    Publisher Summary

    Symbols

    1 Prevention or Delay of Separation of Flow

    2 Flow Control by Provoking Separation

    Work Sheets

    Author Index

    Subject Index

    Copyright

    Pergamon Press Ltd., Headington Hill Hall, Oxford 4 & 5 Fitzroy Square, London W.1

    Pergamon Press (Scotland) Ltd., 2 & 3 Teviot Place, Edinburgh 1

    Pergamon Press Inc., Maxwell House, Fairview Park, Elmsford, New York 10523

    Pergamon of Canada Ltd., 207 Queen’s Quay West, Toronto 1

    Pergamon Press (Aust.) Pty. Ltd., 19a Boundary Street, Rushcutters Bay, N.S.W. 2011, Australia

    Pergamon Press S.A.R.L., 24 rue des Écoles, Paris 5e

    Vieweg & Sohn GmbH, Burgplatz 1, Braunschweig

    Copyright © 1970 Pergamon Press Inc.

    All Rights Reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording or otherwise, without the prior permission of Pergamon Press Inc.

    First Edition 1970

    Library of Congress Catalog Card No. 72-81249

    Printed in Hungary

    08 013441 6

    Preface

    THE problem of the separation of fluid flow is one of those viscous flows which is very important not only for science, but also for practical application. An attempt has been made to compile references up to date mainly in the fields of basic physical processes, analyses, and experiments covering the whole area of interest, giving credit to the original source wherever possible. The content of material has been used for a graduate course at the Catholic University of America, Washington, D.C., in 1962–3, also for a special course at Escuela Tecnica Superior de Ingenieros Aeronauticos, Ciudad Universitaria, Madrid, Spain, in 1964–5, where the author conducted classes as a Fulbright lecturer. Although the material covering the practical application is limited, this monograph is useful for engineers as a reference.

    The unquoted bibliographies have been added in the list of references at the end of each chapter.

    The author wishes to thank the US Air Force, Office of Scientific Research, for grant AF OSR 80-62, 52-63, 52-64, which made it possible to compile this monograph under the guidance of Mr. Milton Rogers, Chief of Mechanics Division, Capt. Hyman H. Album, Capt. Lucius P. Gregg, and Major G. Stalk. The Rev. Everett F. Briggs of Monongah, West Virginia, kindly edited my manuscript. Mr. Richard M. Hartley, of the David Taylor Model Basin, has rendered most valuable assistance in reading my manuscript, in cooperation with Mr. Michael J. Malia of the same institution. Dr. Hyok Sang Lew of the Catholic Universitv of America, and Mrs. Anne A. Ziegler, as well as Mr. Louella A. Chatfield, assisted in arranging and typing the manuscript.

    To these ladies and gentlemen the author wishes to express his gratitude.

    PAUL K. CHANG,     Washington, D.C.

    List of General Symbols

    Subscripts

    CHAPTER I

    Introduction to the Problems of Flow Separation

    Publisher Summary

    This chapter provides an overview of the problems of flow separation. The problem of flow separation is, perhaps, one of the most important hydrodynamic problems to be investigated intensively to find its solution satisfactorily. Because of the complexity of the problem, a rigorous definition of flow separation and stall should be made. The classical concept of flow separation is because of viscosity; therefore, it is often expressed as boundary layer flow separation or boundary layer separation. The chapter highlights classical concept of flow separation is outlined and then the generalized definition of flow separation. The problem of flow separation is as old as that of boundary layer theory. Ludwig Prandtl, the father of the boundary layer concept, was concerned about flow separation before he started his work on the boundary layer. The chapter highlights multiple instabilities of the overall flow field and of various subfields. As Reynolds number increases, the resulting unsteady, three-dimensional, and interacting vortical patterns become keys to some of the perplexing experimental observations reported by various investigators.

    Symbols

    Subscripts

    Superscripts

    The subject of fluid flow separation is one of the many aspects of viscous flow, which is very important but complicated. Because of flow separation, energy is lost. In cases of external flow at subsonic speeds, such as in airborne vehicles, the stream line deviates, the drag increases, the lift decreases, and reverse flow and stalling occur. In the transonic speed range, control and structure problems are created by flow separations. For cases of internal flow, separation can cause reduction in efficiencies. The optimum performance of fluid handling devices such as fans, turbines, pumps, compressors, etc., can only be predicted with accurate understanding of flow separation, because the separation occurs just prior to or at maximum loading. The successful operation of the simplest and most common devices can depend also upon flow separation, as, for example, the throttling action in household water faucets.

    Flow separation can also be useful in engineering applications. For example, a thin airfoil which is suitable for high-speed flight may be made suitable for low speeds by separation of the flow. If the flow is allowed to separate over a portion of the upper surface and then reattaches and remains attached, a very thick pseudo-airfoil results. This thick airfoil is better suited for low-speed flight.

    Another example of desirable flow separation is that caused by placing a spike in front of a blunt body traveling at supersonic speed. Because of the presence of the spike, the flow may separate on the spike and form a cone-shaped flow region in front of the blunt nose. Because of this conical separated flow region, the shock wave will be changed from one nearly normal to an oblique one, which reduces the head drag considerably. Escape capsules and other stores, which must be ejected and recovered from high-speed vehicles, improve their performance by utilizing flow separation.

    Components of high Mach number vehicles, engines, nuclear reactors, reentry vehicles, etc., operate in high-temperature environments. Therefore, problems of heat transfer are incorporated into the problems of hydro-and aero-dynamics. Then, heat transfer problems and flow separation problems, each sufficiently difficult separately, must be considered in combination.

    As a problem area, flow separation has been worked on by many scientists, but much work still remains to be undertaken in this portion of the field of fluid mechanics.

    1 Mechanics of Flow Separation

    The problem of flow separation is, perhaps, one of the most important hydrodynamic problems to be investigated intensively in order to find its solution satisfactorily. Because of the complexity of the problem, a rigorous definition of flow separation and stall should be made. The classical concept of flow separation is due to viscosity; therefore it is often expressed as boundary layer flow separation or boundary layer separation. In addition, a necessary condition for flow separation is the adverse pressure gradient. The details involving these two factors are discussed later in this chapter.

    In a general sense, according to Maskell’s concept of separation, flow separation is inevitable for flow over the finite dimension. Flow will separate from the solid body surface at the trailing edge and also upstream of it if the required conditions are met there. The flow separation is not only caused by a gradual process, which is the case of flow over a smooth surface, but also by a severe discontinuity of the tangent to the surface (Figs. 1 and 2).

    FIG. 1 Flow separation over a smooth surface

    FIG. 2 Flow separation over discontinuous surfaces

    First, the classical concept of flow separation is outlined; then the generalized definition of flow separation is discussed.

    1.1 Classical concept of onset of flow separation over a smooth curved body surface

    The problem of flow separation is as old as that of boundary layer theory. Ludwig Prandtl, the father of the boundary layer concept, was concerned about flow separation before he started his work on the boundary layer. As a young engineer at the Maschinen Fabrik Augsburg–Nürnberg (MAN), Prandtl found that the computed pressure recovery could not be achieved in actual diffusers. He was occupied for 3 years in figuring out why and how the flow separations and pressure losses were caused. This problem was solved finally by his new theoretical concept of the boundary layer [1]. His concept may be referred to as classical in comparison to the modern development of separation of flow theory.

    The classical concept of flow separation is formed for two-dimensional and axisymmetric steady flow. Prandtl [2] states clearly that the necessary condition for flow separation from the wall is the increasing pressure in the streamwise direction, i.e. positive (or adverse) pressure gradient along the flow path (Fig. 3). This statement holds for compressible as well as incompressible flow. Therefore, it may be said that, in general, flow separation occurs under adverse pressure gradient and with laminar or turbulent viscosity effects. If one of these two factors is missing, then the flow does not separate.

    FIG. 3 Flow in a sharply diverging channel [4]

    For example, by removing the boundary layer the viscosity effect is eliminated. As shown in Fig. 4, Prandtl removed the boundary layer along the wall of a subsonic channel by suction.

    FIG. 4 The boundary layer is sucked away on both walls; the flow is from left to right [4]

    In front of the throat, the pressure decreases in the direction of flow because of the decreasing cross-sectional area of the channel. In this region, the pressure gradient is negative (or favorable); hence, flow adheres completely to the walls. However, behind the throat, because of a large degree of divergency or a sufficiently large degree of adverse pressure gradient, the boundary layer separates from the wall and vortices are formed. But, if the boundary layer downstream of the throat is removed by suction, the flow reattaches to the surface. Föttinger [3] provides other experimental evidence to confirm the two necessary factors which cause flow separation; namely adverse pressure gradient and viscosity.

    Figures 5 and 6 show fluid flowing against two separate walls which were placed perpendicular to the flow direction.

    FIG. 5 Free stagnation flow without separation, as photographed by Föttinger [4]

    FIG. 6 Decelerated stagnation flow with separation, as photographed by Föttinger [4]

    One wall was plain, the other equipped with a thin protruding plate. With the plain wall, there was no separation, but with the protruding plate, the flow separated. The physical phenomena may be explained as follows.

    In front of the stagnation point, a considerable pressure rise occurs along the direction of flow, but flow does not separate because of the absence of wall friction. Near the wall, the flow does not separate either, since the fluid in the boundary layer flows in the direction of decreasing pressure. However, by placing a thin protruding plate, the flow separates, due to the pressure rise in the direction of flow as well as the presence of wall friction [4].

    Görtler and Witting [5] computed the point of separation on the protruding plate by a method of series. The velocity distribution for this corner flow is given by

    where ue(x) is velocity at the outer edge of the boundary layer, L is length of the plate, and u∞ is velocity of undisturbed flow. If n is assumed to be equal to unity, then the point of separation lies at x = 0·126L, but, because the plate is of finite length, n is not equal to unity; thus Föttinger’s picture shows the separation point at about x = 0·15L. The discrepancy between computed value of x with experiment is due, firstly, to the influence of corner flow; secondly, to n ≠ 1.

    The role of viscosity in flow separation is further investigated mathematically by Meksyn [6]. As absolute viscosity μ or kinematic viscosity v tends to zero, and with the corresponding decrease in the thickness of the boundary layer, it is expected that the pattern of viscous flow will approach that of potential flow or flow with no separation. But for laminar flow, if the v is sufficiently small, then the pattern of laminar flow is qualified not by v, but by the shape of the non-dimensional function

    where

    which does not depend explicitly on v; therefore, when v tends to zero, the pattern of viscous flow will not necessarily tend to that of potential flow.

    This finding was made about 20 years ago when a difficulty arose in connection with solving the problem of separation based upon the Falkner–Skan equation and uniformly retarded flow past a semi-infinite plane, because the separation point appeared to be a singularity. With the presence of a boundary layer, the pattern of potential flow is incompatible, since in this case λ(ξ) tends to positive infinity at the rear stagnation point, which is physically impossible. The flow pattern becomes physically possible, however, due to the viscosity which modifies the potential flow. In this case, λ(ξ) does not approach infinity, but reaches a low maximum of about 0·5 and then decreases. Separation occurs a short distance downstream of the maximum.

    Over a slender body if the flow separates close to the rear stagnation point, modification of stream lines of the inviscid flow, because of separation, is quantitatively not large. For a blunt body, however, the separation takes place at a position far from the rear stagnation point and, because of the large deviation of a flow pattern from potential flow, the inviscid flow cannot serve as a first approximation to the viscous flow.

    Within the boundary layer, the effect of viscosity is such that the velocity parallel to the wall changes along the distance perpendicular to the surface, i.e. the velocity gradient ∂u/∂y exists (u is streamwise velocity and y is distance normal to the surface). With the exception of rarefied gas, flow velocity at the wall is zero and, with increasing distance y, u gradually increases and finally reaches ue(where τ is the shear stress), at the point of separation, shear stress is zero or, in other words, viscous force vanishes. At a point downstream of separation, because of the existing adverse pressure gradient there, reverse flow occurs as shown in Fig. 7, and, owing to this reverse flow, the boundary layer thickens considerably. At the point of separation, flow begins to leave the surface at a small angle, maintaining the direction of the adverse pressure gradient.

    FIG. 7 Velocity profile near the separation point

    As seen from Fig. 7, the velocity profile downstream at the separation point has a point of inflection. Schubauer and Spangenberg [7] reasoned that this phenomenon occurs from the relation of the pressure gradient along the wall to the velocity gradient along the normal distance to the wall.

    The steady-state momentum equation of two-dimensional incompressible boundary layer flow is given by

    (1)

    At the wall where u = v = 0, this equation is reduced to

    Now, due to the pressure rise on the wall downstream of the separation point,

    Upstream of separation ∂u/∂y > 0 at the wall. At the edge of the boundary layer where the viscosity effect vanishes and ∂u/∂y = 0, the magnitude of ∂u/∂y decreases and ∂²u/∂y² becomes negative. Hence, at a certain normal distance from the wall, ∂²u/∂y² becomes equal to zero and the velocity profile will have a point of inflection. When the velocity profile has a point of inflection, then flow is unstable and tends to become turbulent.

    In order to compute the point of separation, it is necessary first to find the pressure distribution or streamwise velocity distribution of inviscid flow. As the first approximation, the distribution of static pressure and of streamwise inviscid velocity can be found by neglecting the presence of the boundary layer. Then the exact solution will follow when the boundary layer displace ment thickness is added to the boundary of the solid body, because the displacement thickness defined by

    makes the area A equal to B, so that area (A+C) equals area (B+C). Thus, δ* is the thickness that a layer of fluid (area A+C) would have if it had the same integrated velocity defect as the actual boundary layer (area B+C) replacing the boundary of the body surface by δ*, above which the flow may be regarded as non-viscous. Based upon this new thicker body shape (due to inclusion of displacement thickness), the distribution of pressure and streamwise velocity is determined (Fig. 8).

    FIG. 8 Boundary layer displacement thickness

    The line which starts at the point of separation and connects the points of zero velocity within the viscous layer apart from the wall may be defined as the zero velocity line, as shown in δ being the boundary layer thickness. Mixing of turbulent flow is more pronounced than laminar flow, the value of x for turbulent boundary layer amounting to ten times that of the laminar value.

    The effect of mixing is often expressed as a turbulent shear stress τ = ε(∂u/∂y), and the agency responsible is linked to a viscosity ε and is called eddy viscosity. Since the eddy viscosity ε in the boundary layer possesses a value 100 times larger than the dynamic viscosity μ of the same medium, the turbulent flow has a much greater ability of self-mixing. Therefore it is easy to understand that because of larger mass flux and larger eddy viscosity the turbulent flow is capable of advancing further against a rising pressure compared to laminar flow. Thus the mixing is a very important factor characterizing flow separation which could be expressed in a parameter.

    Schubauer and Spangenberg [7] showed that the boundary layer flow mixing may be numerically illustrated in a ratio of displacement thickness to momentum thickness, H = δ*/θ, where H is the shape factor of boundary layer, and

    is the momentum thickness of boundary layer. Von Karman’s momentum integral equation for two-dimensional incompressible flow is given by

    (2)

    (the dynamic pressure), and subscript w denotes wall.

    When turbulence is introduced, an increase in mixing occurs in a very short distance, so that θ may be regarded unchanged; but the value of δ* depends on the amount of fluid through which the retardation is distributed, so the mixing affects δ* directly. The values of δ* change abruptly if the laminar boundary layer becomes turbulent. This can be easily seen from the definition of δ* and Fig. 9. Because ∂u/∂y of the turbulent boundary layer is much larger than that in laminar flow near the wall, δ*/δ is much less than that of laminar flow. Now the numerical values of H may be illustrated for the flat plate flow with zero pressure gradient.

    FIG. 9 Effect of mixing on δ* and θ [9]

    For this case, H for laminar flow is about 2·6 and descends to 1·3 if flow becomes turbulent. Due to mixing of turbulence, δ* is decreased to half of its original value. Since the value of H of turbulent flow is much lower com pared to the value for laminar flow, it may be stated that if rate of increased mixing is high, then the value of H is reduced. But by computing the von Karman equation (2), H can be reduced by decreasing the pressure gradient, and it can be expected that decreased pressure gradient and increased rate of mixing will have similar effects on the boundary layer characteristics [7].

    1.2 Examples of subsonic flow separation

    The flow separation on the body surface is discussed in several classical examples.

    1.2.1 AIRFOIL

    An airfoil yields optimum aerodynamic characteristics of high lift and low drag if the flow is attached to the surface, as shown in Fig. 10. However, if the airfoil is under a sufficiently large angle of attack, then the flow on the upper surface separates as indicated in Fig. 11, and the stream pattern over the upper surface is far apart from the optimum design condition. Furthermore, a considerable size of separated region is created with the formation of vortices. This kind of separation on an airfoil is called stalling, expressing the unfavorable aspect of engineering application.

    FIG. 10 Attached flow over an airfoil [10]

    FIG. 11 Viscous effects in flow around an airfoil (positive angle of attack) [11]

    The stalling is closely connected with closed separation bubbles, which are due to dynamical characteristics of flow setting up circulating motion on the airfoil surface. Depending upon the length, the closed bubbles are short or long. The short bubble is enclosed between the points of separation and reattachment, and its length is the order of 1 percent of chord length, which does not significantly influence the pressure distribution. But if the short bubble of a laminar layer breaks down on the airfoil, then leading edge stall occurs and causes sudden increase of drag and loss of lift, as shown in Fig. 12.

    FIG. 12 Lift and drag of an airfoil [12]

    With a long bubble whose length is about 2 or 3 percent of chord, pressure distribution slightly changes, but its breakdown does not lead to a complete separation of flow; instead, the separated flow passes over the body surface and reattaches further downstream.

    1.2.2 FLOW SEPARATION AROUND A CYLINDER AND A SPHERE

    The physical phenomena of flow separation around a cylinder and a sphere depend on Reynolds number. A classical investigation of flow separation around a circular cylinder or a sphere may serve to illustrate the dependence of flow separation on Reynolds number.

    If the frictionless and incompressible flow theory is applied, then the stream lines drawn by the solid curves result, as shown in Fig. 13. Because of the absence of viscosity assumed by the potential flow theory, flow does not separate and adheres to the body surface. In very slow motion or so-called creeping motion, in the region of Red = u∞d/v ≤ 1, the inertia force is negligible in comparison to viscous force. The resistance caused by such motion is due to the forces necessary for the deformation of the viscous fluid particles. For very small Reynolds numbers such deforming action extends to a large distance from the body compared to the small limited boundary layer thickness at large Reynolds number.

    FIG. 13 Separation of flow around a circular cylinder

    The pressure field in slow motion satisfies the potential equation, and no separation occurs. In this case, heat transfer takes place by conduction alone. By increasing the Reynolds number, the flow separates from the body and the flow pattern may be indicated by the broken curves. Flow is attached to the surface from the stagnation point back to at least point A for all Reynolds numbers. This point A is designated as the point of laminar flow separation. The perfect fluid theory predicts the tangential velocity distribution outside the boundary layer by the following equations:

    These formulas indicate flow acceleration up to φ = 90°, but the laminar flow has already separated at φ = 80–85° on the surface of a cylinder in a region of favorable pressure gradient of potential flow.

    The reason for this discrepancy is due to the fact that the ideal fluid theory predicts satisfactorily only near the front stagnation point, whereas actually the maximum velocity is reached at about φ = 70° instead of φ = 90°, as predicted by the ideal fluid theory, and its value is equal to 1·6 u∞ [6]. Thus, due to the flow separation, the point of the maximum velocity shifts upstream to φ = 70°. Downstream of φ = 70° the static pressure rises, and at φ > 80°, the flow separates if the flow is laminar.

    If the flow is turbulent, then, as mentioned previously, momentum exchange of turbulence takes place more vigorously and, because of such an exchange, the turbulent flow is more capable of resisting the adverse pressure gradient and friction, and flow separates further downstream compared to the laminar flow. Thus the point of separation of the turbulent flow around a circular cylinder is located at point B110°.

    Behind the separation point, a wake is formed where the reverse flow occurs and vortices are formed. Because the laminar separation point is located further upstream, compared to turbulent separation, the size of the wake with laminar separation is larger in comparison to the wake of turbulent separation, and a larger total drag is expected with laminar separation in comparison to turbulent flow separation.

    The streamline patterns as a function of Red are shown in Fig. 14.

    FIG. 14 Types of flow around an infinitely long cylinder [13]. A, at Reynolds numbers less than 1. B, at Reynolds number of 20. S is the separation point of the laminar boundary layer. C, at Reynolds number of 174. One stage of the flow which varies with the time as eddies are shed alternately. Traverse with total head tube along AA′ shown at the right, with width of free laminar layer indicated. D, boundaries of free layer at Reynolds numbers of 5000 and 14,480, and graphs of variation of width of free layer with distance X downstream. D is the diameter of the cylinder. The solid curve in the graph is for a Reynolds number of 5000; the dotted curve is for a larger Reynolds number or for a stream of greater turbulence at a Reynolds number of 5000. E, location of separation of laminar boundary layer and general character of flow at Reynolds number of 80,000. Transition occurs simultaneously with separation. F, location of separation of turbulent boundary layer and general character of flow at Reynolds number 1,000,000. Transition occurs in the laminar boundary layer before separation

    At a Reynolds number of ∼ 10, a laminar flow separation occurs and two weak eddies form in the rear of the cylinder. Due to the flow separation, pressure drag increases and amounts to about one-half of the total drag.

    At Reynolds numbers on the order of 100, flow varies with time, and vortices separate alternately. The region of wake stretches out, and von Karman’s vortex street forms. Since the wake region is fairly large, pressure drag predominates.

    At Reynolds number 5000, the ratio of the width of a free layer with respect to the diameter is small in downstream direction in comparison with the ratio at a Reynolds number about 15,000.

    In a range of Reynolds numbers of 10³–10⁵, the flow separation is of a laminar nature but the wake is turbulent. The drag coefficient remains approximately constant, as seen in Fig. 15, which shows CD as a function of Red and the total drag is mostly pressure drag. The location of the laminar separation point is at φ = 80–85°.

    FIG. 15 Drag coefficient vs. Reynolds number for long circular cylinders and spheres in crossflow [14]

    At Reynolds numbers larger than 10⁵, the flow in the boundary layer becomes turbulent and the separation point moves downstream and locates at about 110°. The wake region reduces greatly in comparison with the laminar separation, and, correspondingly, the pressure drag also reduces, as seen in region of Red = 500,000 in Fig. 15 [13, 14].

    In decreases with increase of stream turbulence.

    At the beginning of the twentieth century, Prandtl and Eiffel measured resistance of spheres. Prandtl measured a drag coefficient of CD = 0·44, while Eiffel measured CD 90°, and when a wire ring was fastened at a for ward position on the sphere, the flow adhered further downstream, and the measured value of CD for this flow condition was 0·176. Hence, it is clear that Prandtl measured at first CD with laminar flow separation yielding CD = as 0·44, while Eiffel measured CD = 0·176 with turbulent flow separation [15]. The measurements of static pressure distribution around a circular cylinder are very useful for understanding the flow separation.

    As Fig. 16 shows, there exists a considerable difference between the measured pressure and predicted pressure of the potential flow theory. It is also shown that the pressure distribution is affected by the Reynolds number.

    FIG. 16 Pressure distribution around a circular cylinder [4]

    At subcritical Reynolds numbers, the flow separates at the point of laminar separation. At about φ = 70°, the pressure rises, as mentioned previously, because the maximum tangential velocity occurs at about φ = 70° reaching its value of 1·6×u∞. This rise of pressure causes the flow separation at φ = 82°. In the separated region, the static pressure remains almost constant in the region of φ = 130–230°.

    At supercritical Reynolds numbers, the flow separates at the point of turbulent flow separation at φ = 110° and the rise of the static pressure starts at about φ = 90°. The magnitude of adverse pressure gradient at supercritical Reynolds numbers in the region of φ = 110–120° is very large, and static pressure reaches a much higher value at φ = 130° compared with the static pressure of the laminar separated flow in the same region. Similar to the laminar separation flow region, the static pressure within the turbulent separated flow region remains almost constant. In the region of the stagnation point, static pressure measured by experiment and predicted by the potential flow theory is equal to others in the past. On the whole, the pressure distribution at supercritical Reynolds numbers differs less from the theoretical prediction than at subcritical Reynolds numbers. As a result, the total drag of a circular cylinder at large Reynolds numbers is less than at small Reynolds numbers. This fact is seen in Fig. 16 by noting the deviation of the real pressure distribution from the ideal.

    Recently, Morkovin [16] summarized and interpreted the new experimental and theoretical studies related to the flow around circular cylinders across an extended Reynolds number range, emphasizing features of the generation of attached and free vorticity and their dynamics. The multiple instabilities of the overall flow field and of various subfields is also carefully studied. It appears that as Reynolds number increases, the resulting unsteady, three-dimensional, interacting vortical patterns become keys to some of the perplexing experimental observations reported by various investigators.

    Figure 17 shows the pressure distribution around a sphere. For a sphere at supercritical Reynolds numbers, the static pressure in the turbulent separated region is not constant but increases in the region of φ = 140–220°, in contrast to the laminar separated flow region where the static pressure remains almost constant.

    FIG. 17 Pressure distribution around a sphere [4]

    It will also be noticed that the static pressure of the laminar region of a sphere is almost constant, but that the pressure coefficient level is three times higher than the coefficient within the same angular region of a circular cylinder.

    1.3 Wake flow at high speeds

    Recently, the problem of wake flow at high speeds has been investigated extensively. Because of its long trail, the wake is a significant source of observation, i.e. electron and radiating species due to high temperature. For example, a meteor trail may reach 15 miles at an altitude of about 100 miles. At a flight velocity of 40,000 ft/sec at the outer part of the wake, temperature may reach 6500°K at 50–100 diameters behind the body [17]. A typical turbulent wake behind a blunt body, an intermediate body, and a slender body at hypersonic speeds is shown in Figs. 18 and 20.

    FIG. 18 Wakes behind bodies at high supersonic speeds. (a) Cylinder at Mach number 5·8 [18]. (b) Blunt cone at Mach number 3·18. Red = 3·58 × 10⁶; p∞ = 1 atm. (Courtesy of US Naval Ordnance Laboratory, White Oak, Maryland)

    FIG. 20 Wake behind slender body at hypersonic speeds. (Courtesy of US Naval Ordnance Laboratory, White Oak, Maryland)

    Furthermore, the velocity and enthalpy profiles just upstream of the neck and just downstream of the neck are shown in Fig. 19.

    FIG. 19 Typical inner wake velocity and enthalpy profiles just upstream and downstream of neck [17]

    The wake at high speeds may be laminar or turbulent, and the transition Reynolds number based upon the local flow conditions and distance x′ from the neck to transition location is, for a blunt body,

    For a sharp body, Retr based upon diameter and local flow condition is about four times as large compared to that for a blunt body [19].

    The hot outer wake behind the blunt body is formed by gas passing through the nearly normal portion of bow shock which is compressed and heated in the shock layer. The inner wake is formed by turbulence which originates in the flow region with highest velocity gradient and by the coalescence of the free shear layer shed from the body surface when the boundary layer leaves the surface. The zero velocity line, on which the tangential flow velocity is zero, originates at the separation point on the body, and is turned back at the neck because the pressure rises by the flow deflection. The fluid above the zero velocity line flows to form the inner wake. Near the neck, the turbulence is confined to a narrow region around the wake axis spreading out rapidly by feeding on the surrounding gas. All of the stream lines originally in outer wake are engulfed downstream by the turbulent wake. Behind the slender body, the wake is cold. In contrast to the hot wake which is primarily associated with bow shock wave, the cold wake may be considered entirely due to the viscous effects associated with the boundary layer and base mixing. Hypersonic wake characteristics, as compared to those at low speeds are given by Lees and Hromas [17] as follows:

    (a) Larger scale vortex does not form even with laminar flow.

    (b) Laminar free shear layer is much more stable at hypersonic speeds.

    (c) Initial value of the momentum defect or drag contained in the inner wake is one or two orders of magnitude smaller than the total body drag.

    But the following characteristics are similar at hypersonic and subsonic speeds as outlined by Fay and Goldberg [20]:

    (1) Well-defined wake disturbance frequency.

    (2) Form of Strouhal number variation with Reynolds number.

    (3) Coincidence of onset of vortex-shedding with onset of wake turbulence.

    (4) Correlation of decay of large-scale vortex system into smaller scale turbulence with Reynolds number.

    (5) Double-thread laminar wakes.

    (6) Non-axial symmetry.

    (7) Wake growth.

    The hot outer wake is cooled mainly by turbulent diffusion and conduction; but under equilibrium conditions, the outer region remains hot even after isentropic expansion to ambient pressure. The inner wake or the turbulent core, which has undergone viscous loss both around the body and in the free shear layer, can also be hot. The turbulent core is cooled by its growth and turbulent mixing, but it requires hundreds of body diameters downstream to do this because it is surrounded by the hot outer viscous region.

    The turbulent wake growth downstream of the blunt body may be given by r x¹/³, where r is wake radius in calibers and x the distance measured in calibers behind the body [21].

    Growth of a cold wake is slow, and it cools much faster than that of a blunt body. But sufficiently far downstream, wakes of all bodies are unique functions of the product of total drag coefficient and base area: CDA.

    1.4 Flow separation by shock interaction

    When the flow velocity over the body is high and becomes supersonic, a shock wave is formed and the shock wave interacts with the boundary layer. Because of this interaction, flow may separate from the surface.

    Similar to the case of subsonic flow, flow separation is caused by the pressure rise in the direction of flow. However, in this case, the pressure rise is caused by a shock wave intersecting flow along a surface. For an understanding of the physics of flow separation due to shock wave-boundary layer interaction, experimental observation and evaluation of measured data are cited here.

    Lange [22] summarized the two-dimensional shock interaction experimental results on flat plates (available up to July 1953). Models of this experiment were flat plates on which steps and wedges were placed. An incident shock wave impinged upon the plates so that sufficiently large adverse pressure gradient was created to cause flow separation. Across the shock the pressure rises, and the pressure difference across the shock is spread out in the lower level of the boundary layer. Thus, the pressure gradient appearing at the wall is determined by the properties of the boundary layer and by the strength of the shock wave. However, flow separation is governed mainly by the pressure rise through the shock wave, hence it is possible that there exists a critical pressure rise which is just strong enough to cause the flow separation. It is convenient to investigate the flow separation by interaction of shock with the laminar boundary layer separate and distinct from the flow separation with the turbulent boundary layer.

    For the laminar boundary layer, Liepmann et al. [23] conducted tests with flat plates. The results show the following trends: the Reynolds number effect on the critical pressure appears to follow the inverse square root of Reynolds number (the Reynolds number definition is based upon the distance from the leading edge of the plate to the point of intersection of the shock wave and the boundary layer); the critical pressure coefficient decreases with increasing Mach number (Fig. 21). These trends of Reynolds number and Mach number agree with the predictions of the following equation:

    FIG. 21 Critical pressure coefficients for laminar boundary layers [22]

    For the turbulent boundary layer, experimental data are discussed by using the following two methods: (a) the wedge technique, and (b) the incident-shock technique.

    FIG. 22 Separation of turbulent boundary layer by step. Obtained in Langley gas dynamics laboratory [22]

    The measurement of pressure distribution in the separated region produced by the wedge and interacting shock shows that, for wedge angles greater than a certain value, the pressure distribution has an inflection point, and the value of the measured Δp/q∞ at the inflection point remains almost constant with further increases in the wedge angle (Fig. 23).

    FIG. 23 Separation of turbulent boundary layer by wedge. Obtained in Langley gas dynamics laboratory [22]

    The measured value of Δp/q∞ at the inflection point is essentially constant and independent of Reynolds numbers in the range from 12 × 10⁶ to 32 × 10⁶ and at M∞ = 3·03.

    Figure 24 shows the local flow separation at an incident shock wave.

    FIG. 24 Incident shock interaction with boundary layer [22]

    From Fig. 24, it is seen that the pressure rise across the incident shock wave is produced upstream in the boundary layer, and, as a result, the boundary layer thickens rapidly. Furthermore, a fan consisting of compression waves is formed immediately in front of the intersection of the incident shock wave and the boundary layer.

    The physical process of shock interaction is as follows: when the incident shock hits the boundary layer it is reflected from the boundary layer as an expansion, and this shock-expansion process turns the flow toward the plate; thus the boundary-layer thickness growth is reduced. A second set of compression waves is formed behind the point where the incident shock hits the boundary layer, and because of the compression wave, flow direction becomes parallel to the plate surface. Thus the separated region may be very small. The pressure distribution varies along the plate with the shock-wave formation. Along the plate, the pressure rises up to the point of separation of flow, and, after this point, magnitude of pressure gradient reduces until the second compression fan of waves increases the pressure to the highest value. Therefore, there is an inflection point in the pressure distribution. A similar observation is made for the flow over the wedge.

    In flow separation due to shock interaction with a turbulent boundary layer, pressure difference varies with Mach number as shown in Fig. 25.

    FIG. 25 Mach number effect on peak pressure coefficient for turbulent boundary layer [22]

    These final results of turbulent flow are similar to those measured for laminar boundary layer; however, between the laminar and turbulent boundary layers there is an appreciable difference in the magnitudes of the upstream influence of the pressure rise across the incident shock waves and of the pressure coefficients.

    The characteristics of shock interaction with the boundary layer phenomena are summarized as follows: for supersonic flow over the flat plates, if the flow is laminar, the pressure coefficient for separation depends on Reynolds number; but if the flow is turbulent, the pressure coefficient depends very little on Reynolds number, if at all. However, the pressure coefficient depends on Mach number whether flow is laminar or turbulent. For a supersonic flow in a shock region, there exists a pressure gradient normal to the wall of considerable magnitude contrary to the subsonic flow. For a subsonic boundary layer, the pressure gradient normal to the wall is so small that the static pressure may be assumed to be constant across the boundary layer.

    Classifying the whole interaction phenomena with a boundary layer, Young [24] distinguished the following five cases:

    (1) Laminar boundary layer in front of shock remains to exist as laminar beyond the shock, with no flow separation accomplished. This case occurs for a weak shock when the Mach number of undisturbed flow is slightly larger than unity. Due to such a weak shock, the pressure rise across the shock is small.

    (2) Laminar flow in front of shock separates because of adverse pressure gradient and reattaches to the surface. This has been observed with Rex = 8·45 × 10⁵.

    (3) Laminar flow separates completely in front of shock, but flow does not reattach to the body surface. For this case, shock is oblique in form.

    (4) Turbulent flow in front of a shock remains attached to the body surface behind the shock wave. For this case, the shock was normal, and no flow separation occurs.

    (5) Turbulent flow attached in front of a shock wave separates behind the shock. For this case, the shock is oblique.

    In general, turbulent flow is more capable of resisting separation due to interaction, compared to laminar flow. Thus flow separation can be suppressed more easily in low as well as in high-speed flow regions if laminar flow becomes turbulent.

    A flow separation is also caused by dust particles in supersonic flow. Brinich [25] observed a separated flow region ahead of a blunt body at a free stream Mach number 5. Such a separation was caused by the impingement and subsequent upstream reflection of dust particles present in the shock-tube gas. This observation is quite significant because the static pressure gradient of free gas stream over the body was negative and, despite the negative pressure gradient, flow separated. When a separated region is formed on the nose, i.e. ahead of the blunt body, then the transition point on the after-body is moved upstream and subsequently the extent of the laminar flow on the after-body is considerably reduced.

    The position of separation is affected by heat transfer. This subject has been investigated by Gadd [26]. At high speed flight, due to aerodynamic heating, heat flows from the hot gas to cool body surface. And, if the hot body is immersed in the cool gas stream, then heat flows from body surface to gas. The heat transfer affects the pressure gradient and the extent of separated region and, consequently, the position of separation. By cooling the wall, pressure gradient deepens and the extent of separated region is reduced, delaying the separation.

    A simple model of flow over a wedge at supersonic speed as shown in Fig. 26 was introduced by Gadd [26] to explain this phenomenon physically.

    FIG. 26 The flow up a wedge on the wall [26]

    S and R denote the points of separation and reattachment respectively. The growth of boundary layer depends upon the strength of adverse pressure gradient acting on the boundary layer, and the pressure distribution is a result of the simple wave compression due to boundary-layer thickening. The heat transfer influences the equilibrium between two processes. By cooling the wall upstream of interaction, although the pressure is constant, the velocity profile becomes fuller and the boundary layer thinner, as shown in Fig. 27.

    FIG. 27 The effect of cooling on the shape of the upstream profiles [26]

    Since the separation is caused by the insufficient momentum near the wall, behavior of density and velocity may clarify the effect of heat transfer on separation. For the cooled wall, density of gas becomes larger, whereas the viscosity becomes smaller compared to no heat transfer. Thus, increase of density coupled with a greater velocity increases the momentum and delays the separation. The opposite is also true. By heating the wall, separation takes place upstream compared to no heat transfer.

    2 Modern Development of Flow Separation Theory

    Recently, in addition to the boundary-layer concept, the stability of the overall flow pattern, particularly as affected by mixing between main stream and dissipative layers, has been emphasized in the attempt to solve the stall characteristics. Considerable accumulation of stagnant fluid occurs in a stall. More details of stall are given in later parts of this chapter.

    Cochran and Kline [27] observed four flow regimes in simple, plane-wall, two-dimensional diffusers as the included angles varied:

    (1) A regime of well-behaved unseparated flow.

    (2) A regime of large transitory stall in which the separation varies in size, intensity, and location with time.

    (3) A regime of essentially two-dimensional, relatively steady, fully developed stall in which the fluid flows with little or no expansion.

    (4) A regime of jet flow in which flow separates and proceeds in a fashion similar to that of jet.

    Cochran and Kline observed the jet-like flow of separation which has been assumed by Crocco and Lees [8], but Crocco–Lees’ theory is concerned only with steady flow based upon this jet-like flow; therefore transitory stall of item (2) is out of the scope of this theory.

    For future development of completely reliable prediction of stall, additional studies are necessary not only of the mean flow, but also of transient characteristics of both the boundary layer and the stall [28].

    The classical concept of flow separation can not be extended adequately to three-dimensional flow separation. Therefore a new approach should be made to deal with the flow separation on an arbitrary shape of body.

    2.1 Three-dimensional separation: generalized concept of flow separation

    The classical definition of two-dimensional and axisymmetric flow at zero incidence identifies separation of flow by the inception of reverse flow, the separation point being the forward boundary of a vortex sheet embedded within a separated region. But this concept is not applicable for three-dimensional flow separation, as Moore [29] and Hayes [30] found. If a reattachment follows after separation, then the flow pattern will be as Fig. 28 indicates. In this case, the vortex sheet is completely embedded in a thin boundary layer.

    FIG. 28 Reattachment of flow [29]

    Moore [29] states that a separated region on a three-dimensional body consists of a vortex sheet embedded between the body surface and a stream surface attached to the body in a closed curve, as shown in Fig. 29. The arrows indicate possible direction of resultant shear stress at the surface and the outside of the separated region. The physical consideration where and when the flow separates is not always clear for the three-dimensional flow. Therefore Moore proposed that a separated flow region be regarded as a vortex sheet embedded in the boundary layer, remaining flat against the body if the assumption of a thin boundary layer is valid throughout the region.

    FIG. 29 Top view of three-dimensional separation [29]

    2.1.1 SURFACE STREAM LINE

    In a three-dimensional flow, separation can occur with no reverse flow and zero friction; therefore a more generalized and new approach is needed to describe and define three-dimensional flow separation. This approach is based upon the new concept of surface stream lines.

    Eichelbrenner and Oudart [31, 32] proposed that the criterion of separation in three-dimensional flow should be set up on the condition that the line of separation must be an envelope of the surface stream lines in the solid surface.

    As Eichelbrenner [33] found, the criterion of separation, in two-dimensional flow is given by ∂p/∂z = 0 and by zero friction; but two-dimensional separation is only a degenerate case of general concept of three-dimensional separation [33] with ∂p/∂z ≠ 0, where z is the coordinate in the direction of the crossflow.

    Eichelbrenner was successful in showing clearly the three-dimensional envelope† of the surface stream lines by emission of milk from holes in the surface of a body placed in a water tunnel, as shown in Fig. 30.

    FIG. 30 Envelope of surface stream lines on an ellipsoid [33]

    A clear definition of the surface stream line was made by Maskell [34], who developed an analysis of three-dimensional flow separation based upon the surface stream lines.

    For steady flow, the surface stream lines are tangent at every point to surface flow direction. On the surface, velocity is zero, but there exist stream lines which pass through points (of height h) above the surface; therefore surface stream lines are defined as the flow directions of particles infinitesimally close to the wall. Then surface stream lines are given by

    where x, y, and z are Cartesian coordinates, and u, v, and w are the corresponding velocity components.

    A stream line cannot end in the fluid; therefore stream lines begin at infinity upstream of the body and end at infinity downstream or form a closed curve. Hence, if a surface stream line does not form a closed curve in the solid surface, it must go in the wall at some point of attachment and leave the surface again at some subsequent point of separation S [33].

    There are two types of separation, namely ordinary and singular. Ordinary separation is common in three-dimensional flow; therefore this type of separation is studied first. On the curved surface, both

    are single-valued and continuous, and since ∂u/∂z and ∂v/∂z are proportional to τx and τy, shear stress in the fluid must be single-valued and continuous, and the skin friction lines are the same as the surface stream lines.

    By denoting

    it may be said that α0 is finite, single-valued, and continuous unless both τx and τy become zero simultaneously. Maskell [34] and Dean [35] observed and reasoned that separation occurs at a point S only when two distinct surface stream lines, lying in the solid surface, converge and meet at that point. They then combine and leave the surface in the form of a single separation stream line. At the separation point where two stream lines meet, the values of α0 of both stream lines are the same. This implies that both surface stream lines must be tangent to each other in the plane or the wall, forming a cusp at the separation. Furthermore, these stream lines must be tangent to the wall at the separation point in addition to the formation of tangency between these two stream lines as shown in Figs. 31–34.

    FIG. 31 Ordinary separation [35]

    FIG. 32 Separation with an isolated singular point [35]

    Enjoying the preview?
    Page 1 of 1