Sie sind auf Seite 1von 276

Ernst Heinrich Hirschel

Wilhelm Kordulla

Shear Flow in
Surface-Oriented
Coordinate
Notes on Numerical Fluid Mechanics
Volume 4

Volume 1 Boundary Algorithms for Multidimensionallnviscid


Hyberbolic Flows (Karl Forster, Ed.)
Volume 2 Third GAMM-Conference on Numerical Methods in
Fluid Mechanics (Ernst Heinrich Hirschel, Ed.)
Volume 3 Numerical Methods for the Computation of I n-
viscid Transonic Flows with Shock Waves (Arthur
Rizzi/Henri Viviand, Eds.)
Volume 4 Shear Flow in Surface-Oriented Coordinate (Ernst
Heinrich Hirschel/Wilhelm Kordulla)

Menuscripts should be _II over 1.00 peges. As they will be reproduced foto-
mechanicelly they should be typed with utmost cere on speciel stationary which
will be supplied on request. In print, the size will be reduced linearly to approxi-
metely 75 'It.. Figurll end diagrems should be lenered accordingly so as to produce
letters not smeller then 2 mm in print. The same is valid for handwritten formulae.
Menuscripts lin English) or proposals should be sent to Prof. Dr. K. Forster, Insti-
tut fiir Aerodynemik und Gesdynemik, Pfaffenwaldring 21, 0-7000 Stungan 80.
Ernst Heinrich Hirschel
Wilhelm Kordulla

Shear Flow
in Surface-Oriented
Coordinate
With 40 Figures

Friedr. Vieweg & Sohn Braunschweig/Wlesbaden


CIP-Kurztitelaufnahme der Deutschen Bibliothek

Hirschel. Ernst H.:


Shear flow in surface-oriented coordinate/Ernst
Heinrich Hirschel; Wilhelm Kordulla. - 1. Aufl. -
Braunschweig. Wiesbaden: Vieweg. 1981.
(Notes on numerical fluid mechanics; Vol. 4'

NE: Kordulla. Wilhelm; GT

All rights reserved


© Friedr. Vieweg & Sohn Verlagsgesellschaft mbH. Braunschweig 1981

No part of this publication may be reproduced. stored in a retrieval system or


transmitted. mechanical. photocopying. recording or otherwise. without prior
permission of the copyright holder.

Produced by IDV. Industrie- und Verlagsdruck. Walluf b. Wiesbaden


ISBN 978-3-663-05277-7 ISBN 978-3-663-05276-0 (eBook)
DOI 10.1007/978-3-663-05276-0
Preface

Having worked on problems of three-dimensional boundary-layer flow


for more than a decade, the authors decided to publish some of
their results and ideas as a book. They believe this is worthwhile
to do in spite of the availability of other books on boundary-layer
theory. These books have their emphasis mainly on two-dimensional
flow, on the modeling of turbulent flow, or on prediction techniques.
The present book, on the other hand, is dedicated to the descrip-
tion of three-dimensional shear flow past realistic configurations
and to the provision of formulations for prediction methods. To
this end several sets of governing equations for the treatment of
high-Reynolds number viscous flows are being discussed, together
with approximations of the geometry of general configurations.
Thus the book , which gives a uniform representation of the for-
mulation of the many problems, and not so much a review of the work
done so far at different places, is aimed at the perspective in-
vestigator of three-dimensional viscous-flow problems.

The reader should not allow himself to ~e scared off'by the use of
tensorial concepts and of the index notation, both not too familiar
in classic fluid mechanics. The authors made the experience that
the use of these concepts alleviates much the work on realistic
flow problems. In the book they have emphasized the use of these
concepts and not their mathematical proofs.

The book was written while the first author was head of the Ab-
teilung fUr Theorie und Numerik (Department for Theoretical and
Numerical Fluid Mechanics) at the Institut fUr Angewandte Gasdyna-
mik der Deutschen Forschungs- und Versuchsanstalt fUr Luft- und
Raumfahrt (DFVLR) in K6ln-Porz (Institute for Applied Gasdynamics
of the German Research Establishment for Aerospace Engineering),
and later of the Abteilung fUr Grenzschichten at the new DFVLR-
-Institut fUr Theoretische Str6mungsmechanik (Dep,rtment for
Boundary-Layer Research at the DFVLR-Institute for Theoretical
Fluid Mechanics), G6ttingen. The second author worked at the same
places, and is still with the latter institute.

The authors thank their colleagues from the DFVLR for the many

v
discussions and for the contributions which went into the book.
Those who should be mentioned explicitly are K. ROBERT and his co-
-author R. GRUNDMANN, whose papers form a large part of the basis
of the book, Mrs. V. JAWTUSCH, who performed many computations in
early stages of the work, and D. SCHWAMBORN, who checked through
some of the chapters, and contributed directly to parts of chapter
12. The authors also thank numerous colleagues from outside the
DFVLR, in particular J. P. F. LINDHOUT, NLR for fruitful and en-
couraging discussions. Last not least they thank the ladies who
type-wrote the manuscript and produced the Indian-ink versions of
the figures.

VI
Contents

Preface
Page

1• In troduction .••.••.•.•.•..••.•.•••••••••.•.•••.••...•••

2. First-Order Boundary-Layer Equations ••••••••••••••••••• 10

3. Boundary-Layer Parameters •••••••••••••••••••••••••••••• 19


3.1 Shear Stress •.•.............•....•....•••....... 19
3.2 Heat Flux ••.••..•••.....•.•••..••.•••••••••••..• 20
3.3 Displacement Thicknesses •••••••••••••.•••••••••• 21
3.3.1 Mass-Flow Displacement Thickness •••••••••••••••• 21
3.3.2 Momentum-Flow Displacement Thickness •••••••••••• 23
3.3.3 Energy-Flow Displacement Thickness •••••••••••••• 25

4. Stagnation-Point Solution ••••••••••••••••••••••.••••••• 26


4.1 Boundary-Layer Equations •••••••••••••••••••••••• 26
4.2 Boundary-Layer Parameters •••••••••••••••.••••••• 30
4.3 Start of the Field Computation •••••••••••••••••• 32

5. Quasi-Two-Dimensional Boundary Layers •••••••••••••••••• 34


5.1 Plane-af-Symmetry Flow •••••••••••••••••••••••••• 35
5.2 Infinite-Swept-Wing Flow 42
5.2.1 Leading-Edge Oriented Orthogonal Coordinates •••• 44
5.2.2 Leading-Edge Oriented Non-Orthogonal Coordinates. 48
5.2.3 Locally-Infinite-Swept-Wing Concept ••••••••••••• 52
5.2.4 Initial Data for the Locally-Infinite-Swept-
-Wing Solution .••..••.•.••••••.••••.•..••••..•.. 56

6. Initial Conditions for the Prediction of Boundary Layers


on Wings and Fuselages 62

7. Higher-Order Boundary-Layer Equations •••••••••••••••••• 70

8. Thin-Layer Approximations of the Navier-Stokes Equations 80


8. 1 Introduction ................................... . 80

VII
Page

8.2 Thin-Layer Approximations 91


8.2.1 Boundary-Layer Like Flows 91
8.2.2 Corner-Layer Like Flows •••••••••••••••••••••••• 94
8.3 Integral Properties •••••••••••••••••••••••••••• 96

9. On the Influence of Surface Curvature ••••••••••••••••• 100


9.1 Hierarchy of Governing Equations ••••••••••••••• 100
9.2 Boundary-Layer Flows at the Leading Edge of
Swept Wings ...................................................................... 103
9.2.1 A Criterion for the Validity of First-Order
Theory for Laminar Incompressible Flows •••••••• 103
9.2.2 Results for Zero- and First-Order Theory ••••••• 113
9.3 Some Remarks on General Cases •••••••••••••••••• 116

10. Samples of Surface-Oriented Coordinate Systems •••••••• 117


10.1 Surface-Oriented Wing Coordinate Systems ••••••• 118
10.1.1 General Wing Coordinate Systems •••••••••••••••• 118
10.1.2 Infinite-Swept-Wing Coordinate Systems ••••••••• 128
10.2 Surface-Oriented Body Coordinate Systems ••••••• 129
10.2.1 Pure Cross-Section Coordinate System ••••••••••• 130
10.2.2 Pure Surface-Oriented Coordinate System •••••••• 139
10.3 Coordinate System of BLOTTNER and ELLIS •••••••• 143

11. Conditions of Compatibility at the Body Surface ••••••• 146

12. Local Topology of Three-Dimensional Separation and


Attachment Lines ............................................................................ 156
12.1 General Remarks •••••••••••••••••••••••••••••••• 156
12.2 Local Topology of Separation Lines ••••••••••••• 158
12.3 Local Topology of Attachment Lines ••••••••••••• 172
12.3.1 External Inviscid Flow ••••••••••••••••••••••••• 172
12.3.2 Viscous Flow (Boundary Layer) •••••••••••••••••• 175

13. Similarity-Type Transformation of Boundary-Layer


Equations in Contravariant Form ••••••••••••••••••••••• 180
13.1 General Remarks •••••••••••••••••••••••••••••••• 180

VIII
Page

13.2 General Three-Dimensional Flow •••••••••••••••• 184


13.3 Stagnation-Point Flow ••••••.•••••••••••••••.•• 190
13.4 Plane-af-Symmetry Flow ••••••••••••••••••.•.•••. 193
13.5 Locally-Infinite-Swept-Wing Concept ••••••••••• 195
13.6 Initial Data for the Locally-Infinite-
-swept-Wing solution .••••••••••••••••••••••••• 196

Appendix AI Basic Geometrical Relations ••••••••••••••••••• 199


A.1 General Coordinates ••••••••••••••••••••••••••• 199
A.2 Surface-Oriented Locally Monoclinic Coordinates 200
A.3 Covariant Base Vectors of Surface-Oriented
Coordinate Systems •••••••••••••••••••••••••••• 203
A.4 Transformation of a Vector ••.••••••••••.••..•• 205
A.5 Partial Derivatives 207
A.6 The Covariant Metric Tensor for Surface-
-Oriented Coordinates ••••••••••••••••••••••••• 208
A.7 The Covariant Curvature Tensor baa of the
Surface ••.•••••••••.•••.•.•••••••••••.•.••••.• 210
A.8 The Two Principal Curvatures of the Surface ••• 210
A.9 Directions of the Two Principal Curvatures of
the Surface ••••••••••••••••••••••••••••••••••• 211
A.10 The Contravariant Metric Tensor aaa of the
Surface ••••••••••••••••••••••••••••••••••••••• 211
A.11 The Mixed-Variant Curvature Tensor b a of the
a
Surface ••••.•••••••••••••••..••••••••••.••.•.. 212
A.12 The Metric Tensor gaa and the Curvature
Properties of the Surface ••••••••••••••••••••• 213
A.13 Christoffel Symbols and Metric Factors ••••••••• 215
A.14 Shifters •••••.••••.•••..••.•.•.•••••••••..•... 218

Appendix BI Series Expansion of Vector Quantities in


General Coordinates 221

Appendix CI The Vorticity Vector •••••••••••••••••••••••••• 225

Appendix D: Dimensions and Boundary-Layer Stretching •••••• 228

IX
Page

Appendix E: Calculation of Streamlines, Surface and


Volume Element 236

List of Symbols •..••.••••.••.••••.•..••••••.•.•.•......•..•• 240

References ••••...••.••••.••••..•••.•.•••.•••••.......••••..• 248

Subject Index ..•••.•••••...•.•••••.•...•.•.........•••...••. 259

x
1. Introduction

Viscous shear flows, to which the present book is dedicated, deve-


lop because of the presence of walls due to the no-slip condition
which prevails in general in flows of aerodynamic interest. In
high-Reynolds-number flow viscous effects are important only in
the immediate neighbourhood of the surface of the body considered,
i.e. within the boundary layer. If the boundary layer remains at-
tached to the surface, the interaction with the external inviscid
flow will, in general, be weak. The viscous flow can then be de-
termined by methods based on boundary-layer theory alone, either
on first or second-order theory, on the latter, if e.g. the boun-
dary-layer thickness is large compared with the radius of the lo-
cal surface curvature. At the trailing edge of wings, where two
boundary layers merge to become a free shear flow - the wake -
stronger interactions, both with the external inviscid flow and
with each other occur whose consideration requires quite sophisti-
cated methods. This is even more the case, if the boundary layer
separates from the surface or if strong vortices are present with-
in the boundary-layer flow. Such flows will be predicted in future
by solutions of the Navier-Stokes equations, or by derivates
thereof which allow to consider the pressure field in the whole
flow region in a global manner. Presently inverse methods are be-
ing developed and used to compute two-dimensional flows with small
separation regions, but it is questionable whether these approaches
can be extended to treat three-dimensional flows. Weak interactions
in flows past isolated components, such as wings or fuselages, are
treated currently with combined inviscid-viscous solutions where
the boundary-layer and the inviscid-flow region are computed sepa-
rately as usual, but coupled by means of the displacement surface
or the equivalent inviscid source distribution.

The development of large scientific computing machines during the


past decade supplied the means to solve numerically partial dfffe-
rential equations, in particular the potential, the Euler and the
boundary-layer equations, for practical three-dimensional flows
past wings or bodies. Future computers with increased storage capa-
city and speed are expected to allow to solve derivates of the

- 1 -
Navier-Stokes equations for high Reynolds number flows past isola-
ted components and complex configurations like wing-body-combina-
tions, realistic inlets, turbo-machinery, etc .. In any case, the
results obtained so far are very encouraging. However, in order
to exploit fully the potential of numerical methods in fluid me-
chanics, many questions need to be answered regarding especially
the phenomena of transition and turbulence, and the properties of
general three-dimensional viscous flows with and without separa-
tion. Therefore much experimental, analytical and numerical work
remains to be done, and regarding the first two disciplines, with
more emphasis as before on three-dimensional and unsteady flows.

The objective of this book is to present and to discuss the gover-


ning equations for three-dimensional shear flows close to arbitra-
rily shaped surfaces, and also the approximation of the geometry
of these surfaces. Thus the book is intended to provide some back-
ground for research work in such"flows, for instance analysis of
experiments, provision of solutions, in particular of boundary-
-layer equations, for flow predictions for the investigation of
criteria for the transition laminar-turbulent and for the work on
turbulence models. On the other hand some background for applied
work is being given, for example for the construction of computa-
tion methods in aerodynamics, although no details such as finite-
-difference schemes, are presented. Methods for the computation of
local interaction and combined inviscid-viscous solutions are not
discussed, either. Integration methods are cited, however, in some
chapters and some peculiarities of such methods are pOinted out.
Concerning prediction methods for three-dimensional boundary layer
it is noted that most of the current working methods are either fi-
nite-difference methods, see e.g. [ 20, 42, 49, 52, 58, 62, 67], or
integral methods, see e.g. [ 21, 22, 79, 81]. The major advantage
of finite-difference methods is their, in principle, potential of
high accuracy of prediction since no assumptions, other than those
associated with physical modeling, are involved. Integral methods
require assumptions with respect to velocity profiles and wall
shear stress distributions, exhibit, however, the characteristic
reduction of the number of independent variables by one. There-
fore, for design purposes, the use of the fast integral prediction

- 2 -
methods is nearly indispensible if it has been assured that the
type of flow in question is correctly predicted. Because of the
long computer-time needed, the use of finite-difference methods
is generally accepted for accuracy checks and for basic research
purposes, e.g. investigation of transition properties, turbulence
models, and for the study of means of boundary layer control, such
as suction, blowing, cooling and the like.

The turbulence problem as such is not a topic of this book. The


authors, however, believe with the future design problems in mind
(drag reduction, noise reduction etc.), that in three-dimensional
flows the modeling of the transition laminar-turbulent might pose
larger problems in numerical computations than the modeling of
turbulence for attached flows. It is, in particular, difficult to
understand and describe the many possible transition mechanisms,
see e.g. MORKOVIN [ 64, 651. To date all knowledge about turbu-
lence and its creation is based on the investigation of flows in
wind tunnels, that is at Reynolds numbers which are rather low
compared with free-flight values. This may add additional uncer-
tainties in design processes. Further unresolved problems concern
flow-structure interactions, unsteady effects, and so on.

The book is devoted to steady, three-dimensional, compressible


viscous flows. Important equations are, however, formulated for
unsteady flow. The equations for two-dimensional flows and those
for incompressible flows are, of course, contained in the equations
presented. Care must be taken where thermodynamic effects playa
role since the energy equations are formulated for thermally ideal
gases, and so is the equation of state. All equations are valid
for laminar flows, and they can be applied to turbulent flows, if
the flow variables are considered as time-averaged, and if appro-
priate effective transport properties are introduced. The Euler
equations can easily be derived from the Navier-Stokes equations
given.

One of the major features of this book is that contravariant vec-


tor components are used in the governing equations. This together
with the index notation leads to compact formulations. Some compu-
ter time and storage is expected to be saved when the equations

- 3 -
are integrated in this form, since the transformation of results
back into physical quantities is simple. A further advantage lies
in the fact that the same tensorial concepts are employed for the
geometric approximation of body surfaces and of metric properties.
In particular transformations from and into Cartesian coordinates,
which are often used to compute inviscid flows and in which the
boundary conditions for boundary-layer predictions are thus given,
are very simple (appendix A). Also very simple is the construction
of streamlines (appendix E) .

Cartesian coordinates are used throughout as reference coordinates


for the formulation of the metric properties of the surface-orien-
ted locally monoclinic coordinates which represent another major
feature of this book. All fluid-mechanical equations in this book
are formulated with regard to the surface-oriented coordinates.
They are derived from equations for general coordinate systems
[21 which use a Cartesian reference coordinate system, too.

The surface-oriented locally monoclinic coordinates have the pro-


perty that two families of coordinate lines x 2 = const., x 3 = 0
and x' = const., x 3 = 0 lie in the surface of the body x 3 = o.
These lines are curved and non-orthogonal, in general. The third
family of coordinates x' = const., x 2 = const. is locally ortho-
gonal to the other two coordinates and rectilinear. The x 3-coordi-
nate lines are thus normal to the body surface (locally monocli-
nic), and all metric properties of the entire coordinate system
can be derived from the metric properties of the body surface
alone (appendix A). Locally monoclinic surface-oriented coordi-
nates are best suited for the treatment of flow near body surfaces.
In fact the classic boundary-layer coordinates are a special case
of these coordinates. The coordinates, however, can be used with-
out restriction on convex surfaces only. On concave surfaces the
x 3 -coordinate lines will converge and eventually intersect each
other. As long as boundary-layer theory is valid, however, the
width of the computational domain normal to the surface, i.e. the
boundary-layer thickness, is by definition small compared with the
smallest radius of local curvature of the surface, and in that do-
main no crossing can occur. In spite of the difficulties associa-

- 4 -
ted with concave surfaces, surface-oriented locally monoclinic co-
ordinates are attractive for the use with Navier-Stokes and even
Euler equations because of their simplicity with respect to the
metric properties and in particular because of their boundary-fit-
tingcharacter. Furthermore the Navier-Stokes equations for the two
surface-tangential directions then contain naturally the boundary-
layer diffusion terms, which easily leads to a thin-layer
approximation of the Navier-Stokes equations for high Reynolds
numbers (chapter 8). To circumvent the problems associated with
concave surfaces for global flow predictions, one must replace the
surface-oriented locally monoclinic coordinates away from the sur-
face, where the x 3 -coordinate lines intersect and thus become sin-
gular, by preferably Cartesian or other orthogonal coordinates.

In Fig. 1.1 a possible combination of surface-oriented locally mo-


noclinic with Cartesian coordinates is depicted in a plane x 2 ' =
const. which is a spanwise cross section of a wing. If the profile
had no concave part on the lower side, the coordinate intersection
line A-A could be removed or placed at x 3 ' + -00. A rounded trai-

Fig. 1. 1 Combination of surface-oriented locally monocli-


nic with Cartesian coordinates for a profile.

- 5 -
ling edge would allow to remove the intersection line B-B or to
place it at x 1 ' + 00. Note that grid-stretching or compression is
possible where wanted, see Fig. 1.1. Fig. 1.2 shows another ex-
ample, the wing-body junction in a plane x 1 ' = const .. Here the
location of the boundary of the Cartesian coordinate system de-
pends on the magnitude of the radius r of the fairing between wing
and fuselage. A sharp corner would make the concept inapplicable.
Note, that in practice both coordinate systems must overlap in or-
der to enable the interpolation of data from one mesh to the
other. The need to interpolate is certainly a disadvantage which
is enhanced further on general surfaces where the coordinate lines

X 3'

x ' ::const., x2'=const.


X 1,= const. , x 3'=const.

x ' =const.
vj' x 2=const.
v2'
x ' =const.
x 3=const.

2'
X

Fig. 1.2 Combination of surface-oriented locally mono-


clinic with Cartesian coordinates for a wing-
-body junction.

x a = const. become skewed at x 3 > 0 compared with those on the


surface x 3 = 0 (see appendix A and chapter 7).

Although surface-oriented locally monoclinic coordinates are not


the ultimate solution to the problem of global coordinate systems,

- 6 -
they exhibit definitely advantages: simplicity, boundary fitting
and very easy use on all convex shapes.

Regarding boundary-layer predictions surface-oriented locally mo-


noclinic coordinates include, of course, external-streamline coor-
dinates which are employed by some investigators. Such coordinates
exhibit the advantage of rigorous orthogonality, but are very la-
borious to construct in particular on bodies with non-analytical
surfaces. This holds, however, to a certain degree for any coor-
dinate system on general surfaces, because the metric factors in
the governing flow equations contain first and second derivatives
of the contour function and depend therefore very much on the
smoothness of the data which define the surface. The use of body-
fixed coordinates (e.g. percent-line coordinate systems on wings,
cross-section systems on bodies, see chapter 10), requires that
the metric properties be determined once only except for the re-
gion near attachment points or lines (chapter 6 and 10), whereas
streamline-oriented coordinates have to be computed anew for eve-
ry angle of attack. For this reason only body-fixed coordinates
are considered in this book. The surface is represented by means
of the Gaussian surface parameters. The formulation involves con-
tour angles in order to reduce the abstractness of the metric pro-
perties of the surface.

Following KUX [ 1], ROBERT [ 3] proposes the use of so-called


shifters in his study of higher-order boundary-layer equations.
The shifters, whose use is quite familiar in shell-elasticity
theory, enable one to formulate the governing equations in terms of
the metric properties of the body surface only. Thus the metric
factors for the space above the surface are not described via the
metric tensor of the coordinates in that region, but with appro-
priate terms in the equations, called shifters, which are formu-
lated as a function of the surface metric and the wall-normal dis-
tance (see appendix A.14, and also chapter 7). The shifter con-
cept, however, is used in this book only in chapter 7, especially
to discuss higher order displacement thicknesses because its use
in the governing equations is believed to increase the computatio-
nal effort. It should be noted that on the other hand the use of

- 7 -
shifters eases considerably, for instance, order-of-magnitude con-
siderations, and other discussions of the governing equations.

The main topic of the present book is the tensorial formulation of


the equations for three-dimensional boundary layers. Chapter 2
presents the first-order boundary-layer equations including the
definition of the metric factors. In chapter 3 boundary-layer pa-
rameters are given. A problem hardly recognized in current vis-
cous-flow research is that of integral parameters in three-dimen-
sional boundary layers. The relations for the displacement thick-
ness and the equivalent inviscid source distribution have been de-
rived by MOORE [ 63) and LIGHTHILL [ 561 long ago. The concepts of
momentum- and energy-loss thickness, which are familiar in corre-
lations for two-dimensional flows, are often extended for the use
in integral methods for the prediction of three-dimensional flows.
KUX [ 11, however, shows that these concepts have no physical
meaning in three-dimensional flows. Therefore he proposes momen-
tum- and energy-flow displacement thicknesses, which should re-
ceive more attention, since more and more experiments are being
done in three-dimensional flow, and the correlation of data should
be based on sound physical concepts. In this book only KUX's equa-
tions are given. In chapters 4 and 5 the equations for quasi-one-
and quasi-two-dimensional flows at stagnation points, in planes of
symmetry and on infinite swept wings are being discussed in tenso-
rial formulation. The boundary-layer parameters are given as well.
In conjunction with the problem of providing initial data in rea-
listic aerodynamic design cases characteristic properties of the
boundary-layer equations are briefly reviewed in chapter 6. Be-
cause integration methods are not being discussed in detail, it
is felt that comments on the essential properties such as the do-
main-of-dependence and -influence condition are sufficient.

In chapter 7 higher-order equations for three-dimensional boundary


layers are presented and discussed. A discussion of the Navier-
Stokes equations with respect to high Reynolds number flows fol-
lows in chapter 8. The need for thin-layer approximations is shown
together with appropriate formulations. In chapter 9 it is dis-

- 8 -
cussed in terms of surface curvature where higher-order boundary-
layer theory applies.

The construction of surface-oriented locally monoclinic coordinate


systems on general wings and bodies is being demonstrated in chap-
ter 10. Examples for both general and simple shapes are presented
in much detail, using the contour angle in conjunction with the
surface parameters.

Compatibility conditions at a body's surface are presented in


chapter 11, starting from the Navier-Stokes equations on curved
surfaces and arriving at those for boundary-layer equations. A
contribution to the discussion of the local topology of separation
and attachment lines in three-dimensional flow is made in chapter
12. Chapter 13 finally deals with similarity-type transformations
of the boundary-layer equations in order to reduce the computatio-
nal effort associated with their integration. No final conclusion
has been drawn so far, and the chapter is intended as a stimula-
tion only.

Basic geometrical and tensorial relations needed for the governing


equations in tensorial form, are provided in appendix A. Although
derivations and proofs are not presented, the reader will find all
necessary relations. The discussions focus, however, on surface-
oriented locally monoclinic coordinates. The remaining chapters
of the appendix contain useful formulations and definitions, and,
last not least, the elegant way of computing streamlines with ten-
sorial, e.g. contravariant velocity components.

- 9 -
2. First-Order Boundary-Layer Equations

Consider the surface of an arbitrary body which is defined in the


(x 1 ' , x 2' , x 3' ) Cartesian (right-handed) reference coordinate
system, Fig. 2.1. The boundary-layer coordinate system, a non-
orthogonal curvilinear right-handed coordinate system
(xl, x 2 , x 3 ), is defined on the surface of the body. Although the
lines xl = const. and x 2 const. are defined on t~e surface, the
x 1 _ and x 2 -coordinate need not necessarily be measured along the
surface of the body or wing (for details see chapter 10.). In
general the lines xl = const. and x 2 = const. are non-orthogonal.
The x 3 -coordinate is rectilinear and normal to both, and therefore
normal to the surface, too. Thus the (xl, x 2 , xJ)-system is a
locally monoclinic coordinate system. xl, x 2 - or xa, a = 1, 2 -
are the Gaussian surface parameters.

Fig. 2.1 Orientation of coordinate systems (xl-coordinate lines:


x2 constant, x 2 -coordinate lines: xl = constant).

The surface oriented xi-coordinate system is considered as being


uniquely related to the Cartesian xj'-coordinate system, the

- 10 -
relationships being given by the base vectors a. (x j ') of the
. -l.
xl.-system (see appendix A) .

The boundary-layer equations in curvilinear non-orthogonal coordi-


nate systems have been discussed earlier by several authors (for
instance SQUIRE [80], HANSEN [32]). In a more general way, includ-
ing especially higher-order effects, and using tensorial concepts,
KUX [ 1] recently treated the equations for incompressible flow,
and ROBERT [ 3] those for compressible flow. ROBERT and GRUNDMANN
[ 2] presented the tensorial formulation of the gasdynamic equa-
tions for Newtonian fluids in non-orthogonal and accelerated
coordinate systems.

ROBERT's equations [ 3] are used as starting point for the dis-


cussions in the present paper. The equations are given in dimen-
sionless tensorial formulation.

The first-order boundary-layer equations then read:*)

Continuity equation

(2.1)

Momentum equation for xl-direction

(2.2)

1 1 1 2 1 3 l' 1 2 1 2
P [ V't + v v'l + v v, 2 + v v, 3 + k 11 .( v) + k 1 2v v +

1
k 16 P'l + k 17 P'2 + ( I.L v, 3) ,3

*) Index notation is used to denote derivatives, for example:

1
v 'i -

(see List of Symbols)

- 11 -
Momentum equation for x 2 -direction

(2.3)

p [v 2 ,t + v
12
v'1 + v
22
v'2 + v
32
v'3 +
12
k21 (v) +
12
k 22 v v +

Energy equation (thermally perfect gas)

(2.4)

123
+ v T'1 + v T'2 + v T'3

1 2] [ 12 12
+ Eref ([p't + v P'1 + v P'2 + ~ k 41 (v '3) + k42 v'3 v'3 +

These equations describe the laminar boundary-layer flow on geo-


metries with curved surfaces. They are also valid for turbulent
boundary-layers if the flow variables are considered as time-
averaged, and if the transport coefficients are taken as effective
transport coefficients including the appropriate turbulent
properties. The equations are valid if locally the boundary-layer
thickness 0 is small compared with the smallest radius of cur-
vature Rmin of the surface

(2.5) o « R
min
For more detailed discussions of this and other cases see chapters
7, 8 and 9.

- 12 -
*) 1 2
In equations (2.1) - (2.4) the velocity coordinates v and v
are contravariant velocity coordinates. These are related to
the physical contravariant velocity coordinates v*1 and"v*2 by
means of the diagonal elements a 11 and a 22 of the metric
tensor of the surface 7,12]

(2.6a) a = 1,2.
Ia: (aa) / Lref
*i
The physical contravariant components v are identical with those
used generally in fluid me~hanics and which are there denoted by
vi which, however, is inconsistent with tensor notation. The
v
quantity re f is the reference velocity of the problem. Variables
without bar have been non-dimensionalize'd with appropriate refer-
ence quantities. The lower index e indicates the external in-
viscid flow field.

In chapter 10 it will be shown that for particular cases the co-


ordinates a 11 and a 22 of the metric tensor of the surface are
just the squares of the normalizing lengths Lx1 and L 2 of x *1
*2 x
and x . If a component a" is identical with one, the vector co-
,1.1. *'
ordinate q1. is identical with the physical coordinate q 1.
(see eq. (2.6)). This is always the case in Cartesi~n systems
(where, in addition, contravariant coordinates are identical with
the corresponding covariant ones). Since the third base vector ~3
in the surface-oriented locally monoclinic system is defined to be
normal to both ~1 and ~2' and of unit length, it holds a 33 = 1,
and therefore

(2.6b) v
*3

The velocity component v 3 , parallel to the boundary-layer coor-


dinate x 3 , is stretched here with the square root of the reference
Reynolds number **):

*)often loosely called components


**)The tilde, denoting quantities with boundary-layer stretching,
will be omitted in the following chapters which deal with
boundary-layer equations exclusively.
- 13 -
- 3
(2. 7) ~3 v*
(a 33 " 1).

This stretching is physically motivated and does not, therefore,


affect the geometrical considerations. Because the xi-coordinate
system is locally monoclinic, a 13 and a 23 vanish. Hence the metric
tensor to be considered reduces to

1 2
a11(x ,x)
(2.8) (
a 21 (x 1 ,x 2 ) a 22 (x 1 ,x 2 )

x 1 and x 2 are the Gaussian surface parameters (see appendix A),


x 3 is the stretched normal coordinate*):

(2.9) ;a 1,2;

Lref is the reference length of the problem. laaa relates the xQ-
coordinate lines locally to a physical length x*a, which, however,
in general, is not a coordinate.

The thermodynamic variables static pressure p, density p, static


temperature T, and the specific heat at constant pressure cp(T) ,
as well as the transport properties viscosity ~, heat conductivi-
ty k, and the time tare non-dimensionalized in the following way:

*) The stretching adopted here, with the square root of the


Reynolds number, is the natural stretching for laminar
flows, see also chapter 13, and corresponding, the numeri-
cal prediction is independent of the Reynolds number. This
stretching is, of course, only one possibility, and, parti-
cularly for turbulent flows, scaling of the boundary layer
coordinate with reference to the local displacement thick-
ness may be more efficient. However, then the variation of
the local reference quantity has to be taken into account.

- 14 -
p T
(2.10) p P T
- - 2'
P ref (v ref) Tref

k t
k ref

Reference Reynolds number, Prandtl number, and Eckert number are:

Pref v ref L ref


( 2 • 11 ) pr ref
~ref

Yref is the ratio of reference specific heats C p lev ' and


Mref = vref/aref the reference Mach numger, withr~;ef Efi~ speed
of sound at Tref .

The equation of state is

pT
(2.12 ) P =
Yref(Mref) 2

The metric factors k mn , which are functions of the components


1 2
aas(x ,x ) of the metric tensor of the surface, read:

Continuity equation

(2.13a) [ all a ( a12) 211/2


22 -
where

(2.13b)

is the determinant of the metric tensor of the surface.

xl-momentum equation

(2.14 )

- 15 -
(2.15 )

(2.16 )

(2. 17)

(2.18 )

x 2 -momentum equation

1
(2. 19) k21 2a [a 11
(2 a 12 '1 - all' 2) - a 12 a 11 'l) ,

(2.20) k22 -a all a 22 '1 - a 12 all' 2 ) ,


1
(2.21) k23 2a [all a 22 '2 - a 12 (2 a 12 '2 - a 22 '1») ,

(2.22)

(2.23)

energy equation

(2.24)

(2.25 )

(2.26)

The boundary-layer equations (2.1) - (2.4) are for the incompress-


ible case essentially those of SQUIRE [80). The commonly used
metric coefficients h 1 , h2 and g are related to the components
of the surface tensor aa~ in the following way:

(2.27) g •

- 16 -
The square of the length element on the surface therefore reads

(2.28)

Note that the metric factors simplify considerably if the diago-


nal components of the metric tensor of the surface oriented coor-
dinate system have the properties a 11 = a 11 (x 2 ), a 22 = a 22 (x 1 ).
This is the case, if particular coordinate systems are used (see
chapter 10). In special cases, e.g. for simple geometries, this,
and a 12 = 0, may occur anyway.

Boundary values have to be prescribed at the outer edge of the


boundary layer:

x3
3
(2.29) -+- co (x = 6) :
1 1
v1 ve (x , x 2 ) , v2 = v! (x 1 , x 2 ) ,

T T (x 1 , x 2 ),
e

and at the wall, where no-slip conditions usually hold:

(2.30) 0:

i = 1,2,3
although blowing or suction in the frame of the boundary-layer
assumptions

(2.31) x2,

with v~ 0 (Re~~~2) is possible. The wall temperature may be pre-


scribed:

(2.32) T (x 1 ,

or its first derivative in x 3-direction:

- 17 -
1 2 1 2
(2.33) T'3 (x , x , 0) Tw' 3 (x ,. x )

allowing to prescribe the heat flux qw through the surface.

Initial data have to be prescribed where flow enters the computa-


tional area (see chapter 6)

(2.34)

v
a ( xi ) , T 1 ,2,3 a = 1,2

- 18 -
3. Boundary-Layer Parameters

3.1 Shear Stress

The shear stress tensor has to be considered as a double-contra-


variant quantity [ 2]. Multiplication of its coordinates with the
square root of the appropriate diagonal elements of the covariant
metric tensor yields the physical coordinates of the shear stress
tensor (see chapter 8). Here only the contributions relevant in
zero- or first-order boundary-layer theory are given. In dimen-
sionless form and with the boundary-layer stretching applied to
both, v 3 and x 3 , the physical shear stress component for the
x 1 -direction reads:

(3.1)
T· 13 Lref

(3.1a) - 11 v
.1
'3 ra:t1 1'13

(3.1b) 13 1
l' - 11 v '3

and for the x 2-direction

(3.2)

(3.2a) 23
l'

(3.2b) 2
- 11 v '3

Although the minus sign is not introduced in the definition of


the shear-stress tensor of reference [2], this is done here to
be consistent with Newton's law.

- 19 -
It is interesting to realize that in first-order boundary-layer
theory the influence of surface curvature is not explicitly appa-
rent in the expressions (3.1) and (3.2) for the shear stress
(/a nn is simply a length scale).

In the case of turbulent flows the viscosity in the expressions


(3.1) and (3.2) is being considered to be an effective viscosity,
including the apparent turbulent viscosity, if lowest order tur-
bulence closure is used. This certainly is a crude simplifica-
tion of the turbulence problem, in particular for three-dimension-
al shear layers, but often used because of lack of better and
more efficient models. It poses, however, no principal problem
to incorporate a more sophisticated model into the governing
equations. Equations (3.1) and (3.2), however, are accurate for
computing the shear stresses at an impermeable, hydraulically
smooth surface for both laminar and turbulent flows with no-slip
conditions, since there ~ is the molecular viscosity in both
cases.

3.2 Heat Flux

The physical dimensionless heat flux normal to the surface, after


the usual boundary-layer stretching of x 3 has been introduced, is

-*3 L
q ref
( 3 • 3)
'k ref'T ref ~
ref

The remarks about the use of equations (3.1) and (3.2) in turbu-
lent flows are valid for equation (3.3) in a similar fashion.
While it is sufficient to know the molecular heat conductivity in
order to determine the heat flux at the wall, which is of prime
interest, one introduces an effective heat conductivity, including
the apparent turbulent heat conductivity, away from the surface.
In most cases the effective turbulent heat conductivity is ex-
pressed by means of the effective viscosity and a constant, ef-
fective turbulent Prandtl number.

- 20 -
3.3. Displacement Thicknesses

The remainder of this chapter deals with integral boundary-layer


parameters which often facilitate the correlation of experimental
results, and which may appear in integral solution techniques as
dependent variables. Such integral parameters reflect the influ-
ence of the boundary layer upon the (hypothetical) inviscid flow
past a body, and they, therefore, globally characterize the state
of development of the boundary-layer. The present book follows
the concept of KUX [ 11 who introduces mass-, momentum- and
energy-flow displacement thicknesses. While the concept of mass-
flow displacement thickness in two-dimensional flow has been ex-
tended to three-dimensional flow, MOORE [631, LIGHTHILL [561, it
seems that the concepts of momentum- and energy-loss thickness
used in two-dimensional flow cannot be extended to three-dimension-
al flow, KUX [ 11. KUX proposes to use momentum- and energy-flow
displacement thicknesses, defined in the same way as for the mass
flow. These quantities are shown to appear naturally as dependent
variables in three-dimensional integral boundary-layer equations.
All, moreover, have in common that partial differential equations
have to be integrated, whereas for two-dimensional boundary layers
mere quadratures have to be performed in order to compute the in-
tegral parameters. In this book, however, except for some more de-
tailed discussions in chapter 7, only the results of KUX's inves-
tigation are given.

3.3.1 Mass-Flow Displacement Thickness

The dimensionless mass-flow displacement thickness

(3.4)

is found in zero- or first-order boundary-layer theory in the


present notation by solving the following first-order differen-
tial equation:

- 21 -
(3.5)

This equation is equivalent to MOORE's [63] equation for general


orthogonal surface coordinates. The term at the right-hand side
of equation (3.5) is due to suction or blowing and vanishes for
an impermeable wall. The quantities 0 1 n denote the familiar
x
two-dimensional definition of displacement thickness with respect
to the velocity components va, a = 1,2:

xl-direction

0
61x1 IRe ref
pv
1
(3.6) o1 x 1 - (1 - ) dx 3
L ref J 1
Pe v e
0

x2~direction

6
61x2 IRe ref
pv
2
(3.7) 01 2 - (1 - ) dx 3
x L J 2
Pe v e
ref
0

The lower index e in the above equations denotes the external


(inviscid) flow field, the properties of which are known. Never-
theless, equation (3.5) can only be integrated after the solution
of the boundary-layer equations has been obtained or measurements
have been made, since the evaluation of the integrals (3.6) and
(3.7) requires the knowledge of the velocity and density profiles.

a
For two-dimensional flows --- = 0, and equation (3.5) reduces to
ax 2
the two-dimensional definition of the displacement thickness (3.6),
if v 3 (x 3 = 0) = o.

- 22 -
The mass-flow displacement thickness is often used in inviscid-
viscous interaction calculations to simulate the effective (vis-
cous) shape of a body immersed into some flow, because it re-
presents the displacement of the freestream due to viscosity.
On the new body surface the boundary conditions are then applied
for the inviscid-flow prediction. An alternative is to use the
equivalent wall-outflow velocity on the geometric body as boun-
dary condition reflecting the viscous influence, LIGHTHILL [56],
PIERS et al. [69]. No partial differential equation has to be
solved, but the boundary layer profiles are still needed:

(3.8)
inv

The lower index inv abbreviates inviscid and denotes the sur-
face-normal mass-flow component to be imposed at x 3 = O. Because
Po inv is initially unknown, the use of equation (3.8) may in-
volve some iteration in compressible flows. But since the shape
of the body remains unchanged in the course of an inviscid-vis-
cous coupling procedure, the metric factors of the body remain
unchanged as well, and have to be determined only once if they
are needed for the inviscid-flow prediction.

3.3.2 Momentum-Flow Displacement Thickness

Because the momentum-flow is a vector two momentum-flow displace-


ment thicknesses have to be defined in three-dimensional boundary
layers. The dimensionless thicknesses 6 2 due to the momentum
xQ
flow in the xQ-directions are defined by

(3.9)
62 x Q /Ref
re ,. Q = 1,2 •
Lref

- 23 -
Following the derivation of KUX [ 1] two first-order partial
differential equations result for the determination of 6 2 a in
x
zero- or first-order boundary-layer theory (see also chapter 7):

(3.10) (v 1) 2
[k01 P e e (6 2 1 1 - 6 2x1 )]'1 +
x x

+ [k01 P e v 1 v 2 (6 2 1 2 - 6 2 1)]'2+
e e
x x x

+ (v 1) 2 6 2 1) +
k01 k11 P e e (6 2 1 1 -
x x x

1 2 [ 1
+ k01 k12 P e veve 6 2 1 2 - "2 (6 2 1 + 6 2 2)] +
x x x x

+ (v 2 )2 (6 - 6 2 2) 0
k01 k13 P e e 2 2 2
x x x

and

(3.11) 1 2
[k01 P e v e v e (6 2 1 2 - 6 2 2) ]'1 +
x x x

(v 2 )2 6 22 1]'2 +
+ [k01 P e e (6 2 2 2 -
x x x

+ (v 1) 2 +
k01 k21 P e e (6 2 1 1 - 6
2 1
)
x x x

1 2 1 (6
+ k01 k22 P e v e v e - "2 2 + 6 2 2)] +
[6 2 1 2
x x x1 x

+ (v 2 )2 0
k01 k23 P e e (6 2 2 2 - 6
2 2
)
x x x

with

6
(3.12) 6 a,a 1,2 (no
x xa
2 a J
o sUmmation!) •

- 24 -
Again, because of the boundary-layer integrals (3.12) the momen-
° °
tum-flow displacement thicknesses 2 1 and 2 2 can only be ob-
x x
tained from equation (3.10) and (3.11) after the boundary-layer
equations have been integrated, or density and velocity profiles
have been measured.

3.3.3 Energy-Flow Displacement Thickness

The energy flux is a scalar, and therefore only one equation re-
sults as it is the case for the mass-flow displacement thickness.
Let E be the quantity under consideration: the kinetic energy
per unit mass ~(q)2, that per unit volume ~ p(q)2, or the total
enthalpy h + ~(q)2 And let the dimensionless energy-flow dis-
placement thickness °3 be defined by

(3.13 )

Then, following the concept of KUX [ 11, the form of the partial
differential equation for °
3 is the same as that for the mass-
flow (3.5) without source term if p is replaced by E:

with

(3.15 ) a = 1,2
(no summationl).

Equation (3.5) is a special form of equation (3.14) with E p


and °3 a
x
°1 a·
x

Equations (3.5), (3.10), (3.11) and (3.14) are linear first-order


partial differential equations of hyperbolic type. They can succes-
sively be integrated together with the boundary-layer equations
after the boundary-layer profiles have been obtained.

- 25 -
4. Stagnation - Point Solution

The boundary layer at three-dimensional stagnation points has


been the object of many theoretical investigations. Since a re-
view on the literature available is not intended, only a few re-
ferences are mentioned here [40,53,68]. HOWARTH [40]was one
of the first to present results for a general three-dimensional
stagnation point. He considers the incompressible flow at a nodal
point of attachment. LIBBY [53] reports a solution for compress-
ible boundary layers with heat and mass transfer which is also
capable to handle the flow at saddle points of attachment (for a
de~ailed discussion of nodal and saddle points of attachment see

e.g. LIGHTHILL [57]). Although the above investigations consider


the flow at a stagnation point on a general curved surface,
second-order boundary-layer effects are not taken into consider-
ation. The recent studies of PAPENFUSS [68] are concerned with
second-order boundary layer effects at (nodal) points of attach-
ment with strong suction or blowing.

In the present chapter the boundary-layer equations and parameters


for the steady flow at a three-dimensional stagnation pOint are
given consistent with the governing equations in chapter 2, i.e.
neglecting second-order boundary-layer effects.

4.1 Boundary-Layer Equations

Consider a point S of attachment of flow on a general curved sur-


face, Fig.4.1. Let the origin of the locally monoclinic, surface-
oriented coordinate system (x~) coincide with the stagnation
point S = S (xi = 0) on the surface. As before the, in general non-
orthogonal coordinates x 1 and x 2 are the surface coordinates,
and x 3 is the rectilinear coordinate normal to the surface.

In order to arrive at a quasi-one-dimensional solution which is


needed to start the boundary-layer prediction, it is required that
at least in the vicinity of the stagnation point the coordinates
x 1 and x 2 coincide with external streamlines. Furthermore, these
streamlines must not be too strongly curved close to the stag-

- 26 -
X3
1

X "
Fig.4.1 Nodal point of attachment on a general, curved surface.

nation point so that stream surfaces normal to the body surface


can be defined with sufficient accuracy by the external stream-
lines x 1 and x 2 (see also chapter 5.1). These conditions are
automatically fulfilled if a body and the flow past it exhibits
two planes of symmetry. The assumption of external streamlines
x 1 and x 2 with weak curvature at S assures symmetry of flow .
throughout the boundary layer (in the direction normal to the
wall) at s:

(4.1)
o o

It is then reasonable to assume symmetry for the thermodynamic


variables temperature T and density p and, because of eq.(2.12)
for the pressure p as well:

- 27 -
xl
(4.2)

T'l p , 1 p , 2 o .

If the wall is being heated or cooled equations (4.2) may not


hold. This applies also for the general supersonic flow about an
inclined. blunt body where the point-of-attachment streamline does
not necessarily coincide with the streamline of maximum entropy
behind the curved shock front. However, both equations (4.1)
and (4.2) are required to derive the stagnation point equations.
In fact, instead of eq. (4.2) only the weaker condition of sym-
metric pressure distribution P'l = 0, P'2 = 0 is needed at the
stagnation point. But since symmetric temperature and density
profiles are present at quasi-two-dimensional dividing stream
surfaces (see chapter 5), and since in most cases the stagnation
point is part of such surfaces it is suitable to require condi-
tions eq. ( 4 . 2) .
Using equations (4.1) and (4.2) together with the definition of
a three-dimensional stagnation point:

xl = 0, x 2 = 0, x 3
-> 0:

(4.3) v* 1 = 0, v* 2 = 0,

v *1 , 1 +0, v
*2
'2 +0,

causes the momentum equations(2.2) and (2.3) to vanish. The con-


tinuityequation (2.1) and the energy equation (2.4) don't vanish,
however.

The continuity equation yields

1 2 3
(4.4) p (v ,1 + v , 2) +(pv)'3 o ,

- 28 -
and thus requires the knowledge of the profiles of the gradients
v~l and v~2 in order to be integrated for v 3 as usual. Defining
new dependent variables at xl = 0, x 2 = 0, x 3 ~ 0

Aa Aa (x 3 ) [v*a(X 3 ) ]
(4.5) [v a (x 3 )]'a 'a a = 1 ,2
v'a(aa) (no summation!) ,

the continuity equation (4.4) reads

(4.6) o .

The profiles of the new dependent variables Al and A2 are ob-


tained by integrating the momentum equations (2.2) and (2.3),
after these have been differentiated with respect to xl and x 2
and after the conditions (4.1) to (4.3) have been introduced.

The momentum equations then read:

(4.7) a=1,2,

where the relation v a '3a = va'a3 has been used since it can be
assumed that the va are continuous at S. Furthermore, the pressure
gradient terms have been replaced by the convecitve terms at the
outer edge of the boundary layer.

(4.8)

Note that for orthogonal coordinates the cross derivative of the


pressure is not needed, since k17 = k26 = 0 in that case.

- 29 -
Finally, the energy equation (2.4) reduces to

3
(4.9) cp p v T, 3

The equations given in this section are equivalent to those given


by the authors cited above if the appropriate assumptions and
transformations are introduced. The equations for the two-dimen-
sional and axisymmetric stagnation point are easily obtained
from equations (4.6), (4.7) and (4.9).

4.2 Boundary-Layer Parameters

The tangential shear stresses vanish at the stagnation point


throughout the entire boundary-layer. Their derivatives, however,
are obtained by differentiating the expressions (3.1) and (3.2)
with respect to xl and x 2 :

(4.10) a = 1,2
(no summation!) .

The heat flux normal to the wall, eq. (3.3), is

(4 • 11 )

The partial differential equation (3.5) for the mass-flow dis-


placement thickness reduces to an algebraic expression:

+ A~ 01 2
x
(4.12) +

- 30 -
where 0, 11
x

(4. '3) 0, X
11 a = 1,2
(no summation!).

Similarly, equation (3.8) yields for the equivalent inviscid


wall outflow velocity:

(4. '4)

The partial differential equations (3.10) and (3.1') which deter-


mine the momentum-flow displacement thicknesses 0'x2 and 62x2'
become meaningless at the stagnation point. By differentiating
(3.10) with respect to x' and (3.11) with respect to x2 , one
obtains algebraic expressions for the three-dimensional momentum-
flow displacement thicknesses:

2
62 1 1 + Pi.e 62 , ')
(4. , 5) x x x X""

(4. '6)

with

(4.17) a,B = 1,2


(no summation!).

The energy-flow displacement thickness, finally, is obtained


from equation (3.14):

(4.18)

- 31 -
where

(4.19) U = 1,2
(no summation!),

and E stands for the kind of energy considered.

Note that the expressions (4.12) through (4.19) don't depend


explicitly on the geometry of the body surface near the stagna-
tion point, and that the integral thicknesses are not zero at
the stagnation point.

4.3 Start of the Field Computation

After the stagnation point solution has been found, the dependent
variables in the vicinity of the stagnation point can be computed.
These values can serve as initial values for the field solution.

In Fig.4.2 the vicinity of a stagnation point is shown. The stag-


nation point coordinate system is depicted as the xU-system, and
s
the body (or field) system as the xb-system. The x 3 -coordinate in
each system lies normal to the surface.

2
X
5
=0

,
Xb= const.

X' =0
5
S(X'5 =OX
IS
L =OX 3 ,
, J

Fig.4.2 Stagnation point coordinate system and body coordinate


system (idealized representation) .

- 32 -
For small distances between stagnation point 5 (xs1 = 0, x s2 = 0, x 3 )
and field points P(X~'X~'X3) all scalar quantities p,T,p as well
as v 3 are to first order the same in both points because of the
conditions eq. (4.2).

1 2 3 ) are found follow-


The velocity vector components Vsa in P(xb,xb,x
ing appendix Band eqs. (4. 1), (4.3) and (4.5) :

(4.20) a = 1,2 .

These velocity components now are transformed into velocity com-


ponents vi' (P) of the reference coordinate system in P with help
of eq.(A.20):

(4.21) a~' (P) v j (P) i,j = 1,2,3,


Js s
(v 3 (P) =01).
s

i'
Here ajs(p) are the components of the base vectors ~is of the
stagnation point coordinate system in P(X~'X~).

The final transformation into velocity components V~(P) of the


body coordinate system in P is made with help of eq.(A.24):

(4.22) ai 'b (P) vi' (P) i,j = 1,2,3 ,


(v~(P) =01).

Here af'b are the "inverse components" of the base vectors ~jb
of the body coordinate system in P(X~'X~) following from
eq. (A. 22) •

- 33 -
5. Quasi-Two-Dimensional Boundary Layers

Three-dimensional boundary layers are called quasi-two-dimensional,


if the governing boundary-layer equations explicitly depend on
two coordinates only. Accordingly, the three-dimensional stagna-
tion-point flow, discussed in chapter 4, is called quasi-one-
dimensional because only one coordinate, that normal to the wall,
appears in the governing equations (4.4), (4.7) and (4.9). Quasi-
two-dimensional boundary-layers are present, for example, in
planes of symmetry of flow past, for instance, bodies and infinite
swept cylinders and wings. A well-known example of a plane-of-
symmetry flow is the quasi-one-dimensional flow along the line of
attachment of an infinite swept cylinder or wing [10,16,18]. This
attachment-line flow separates the flow on the upper surface from
that on the lower surface, and may therefore be said to establish
a dividing stream surface. The concept has been extended to finite
attachment lines, then yielding quasi-two-dimensional solutions,
[18,26,38,85,87].

Quasi-two-dimensional boundary-layer solutions represent an essen-


tial ingredient for many three-dimensional boundary-layer predic-
tions, because the finite-difference prediction methods need
initial data on at least one, but mostly on two surfaces (erected
normal to the surface of the body considered, and conveniently
chosen such that their intersections with the body coincide with
lines of coordinates) in order to start the integration of the
field equations (see chapter 6).

If real configurations are considered such as lifting finite wings


or fuselages at an angle of attack, the plane-of-symmetry and
infinite-swept-wing concepts are fairly easy to apply. In most
cases, however, these concepts are only approximately valid. For
instance, the dividing stream surface near the leading edge of
a finite lifting wing, in general, does not represent a plane-of-
symmetry flow [78]. Nevertheless, this quasi-two-dimensional
approach is in many applications sufficiently accurate.

The list of references given in this chapter on quasi-two-dimen-


sional boundary-layer solutions is not complete since no review
is intended.
- 34 -
5.1 Plane-of-Symmetry Flow

A plane-of-symmetry flow exists, in general, only if the geometry


of the body exhibits a plane of symmetry and if the vector of the
oncoming flow lies within that plane. Due to the symmetry a surface-
oriented coordinate system is adopted which is locally orthogonal
in the plane of symmetry (see Fig. 5.1), and which may be non-
orthogonal elsewhere.

X 3'

\
Plane of Symmetry
X 7'
Fig. 5.1 Plane-of-symmetry flow, and (locally) orthogonal
coordinate system in the plane of symmetry at x 2 O.

Such a coordinate system is easily established for a wing with


symmetrical cross sections and an angle of attack of zero degree.
Then the usually employed surface coordinates, given by lines of

- 35 -
constant chord and constant span, are orthogonal along the leading
edge which corresponds to the line of attachment imbedded in a
plane of chordwise symmetrical flow. If, however, there is an angle
of attack the line of attachment moves away from the leading edge
and becomes curved. It must then be assumed that a surface of sym-
metry approximately exists, at least locally: the above chosen
coordinates, however, are non-orthogonal there (see chapter 5.2.4).
In the case of symmetric bodies at angle of attack with a streamwise
plane of symmetry of flow, the surface coordinates usually employed
are orthogonal there.

If the coordinate are orthogonal the metric factors k mn defined


in equations (2.13) to (2.26) reduce considerably because the off-
diagonal term a 12 a 21 of the metric surface tensor vanishes. Sym-
metry-plane flows have been discussed for example by LIBBY et al.
[55], LIBBY [54], WANG [87], HEAD and PRAHLAD [33], CEBECI et al.
(18], CEBECI [17], NASH and PATEL [ 81, HIRSCHEL and SCHWAMBORN
[38],and GRUNDMANN [30].

Plane-of-symmetry flows are characterized. by the variation of the


dependent variables across the plane of symmetry: one variable is
symmetric of the second kind (odd symmetry) while the rest of the
dependent variables is symmetric of the first kind (even symmetry) .
Refering to Fig. 5.1 this means that v*2 is zero in the plane of
symmetry x 2 = 0, but that v*2(x 1 ,x 2 ,x 3 ) = - v*2(X 1 ,_x 2 ,x 3 ) and
therefore v 2 '2 1 0 at x 2 = O. The other variables are finite at
x2 = 0, but their first derivatives with respect to x 2 vanish due
to even symmetry. Since ge·ometrical symmetry is assumed the metric
elements aa~ are even as well. If the coordinate system is ortho-
gonal also in the vicinity of the line of symmetry, a 12 '2 = 0
for x 2 = O. In general, however, this is not true, and a 12 is of
odd symmetry. Taking this into account the metric factors k mn
needed in the following equations of motion, reduce to the follow-
ing at x 2 = 0:

(5.1 ) a

(5.2)

- 36 -
(5.3) k16
all

(5.4) k22 a 22 a 22 '1 ,

(5.5) k27 a 22

(5.6) k41 all

The rest of the metric factors vanishes or will be coefficients


of vanishing terms in the governing boundary-layer equations.
Derivatives with respect to x 2 of vanishing metric factors, how-
ever, remain. Equations (5.1) to (5.6) simplify further, if
all 1 all (x 1 ) (which may be the case, see chapter 10). The set of
governing equations (2.1) to (2.4) reduces for the plane-of-sym-
metry flow, depicted in Fig. 5.1, to (assuming steady flow):

Continuity equation

1 2 3
(5.7) (k O1 p v ), 1 + kO 1 j) A + (k O1 p v ), 3 = 0

where A2 is defined by equation (4.5) applied in the plane of sym-


metry x
2
= 0: A2 = v'2
2 Ix 2 =0 1
= ----
,!a 22
v *2 '2 2
x =0
I•

Momentum equation for x 1 -direction

(5.8) 11 31 12 1
p [v v ' l + v v'3 + k 11 (v) 1 = k16 P'l + (1-1 v '3)'3

2
The momentum equation for the x -direction b~comes meaningless and
has to be differentiated with respect to x 2 in order to yield an
equation for the determination of the quantity A2 remaining in the
continuity equation (5.7). The differentiated momentum equation
(2.3) then reads at x 2 = 0:

- 37 -
(5.9)

2
+ (~L A '3)'3

with k 26 '2

Energy equation

(5.10)

1 1 2
+ Eref [v P'l + J.1 k41 (v '3) J.

Initial and boundary conditions have to be supplied for the inte-


gration of equations (5.7) to (5.10). The initial conditions are
given by the three-dimensional stagnation-point flow (see chapter
4.). The boundary conditions at the wall are usually the no-slip
conditions with prescribed temperature or temperature gradient. At
the "edge" of the boundary layer the inviscid-flow solution provides
the external boundary conditions. Regarding A2 the no-slip con-
dition leads to A2 (x 3 = 0) = O. The external distribution A~(X1)
is often not given or difficult to obtain in terms of velocities.
Then it is necessary to evaluate A2(x 1 ) from the pressure deriv-
3 e
atives. For x + ~, at the edge, equation (5.9) reduces to

(5.11 )

- 38 -
This is an ordinary differential equation for A~ which can be
integrated if the other variables are known. The initial value
for the integration of equation (5.11) is given at the stagnation
point as follows:

2
(5.12 ) 0, x - 0)
0, x 2 _ O.

Since the pressure distribution attains a maximum at the stagna-


tion point expression (5.12) is real (note that k27 < 0). The
sign of A! is positive if there is a nodal point of attachment
(convex surface), and negative for a saddle point of attachment
if the surface is concave with respect to the x 2 -direction. With
the initial value given, equation (5.11) can be integrated by
means of conventional finite-difference techniques. Particular
care has to be taken in choosing the finite-difference represen-
tation such that no singularities occur [38]. This is completely
avoided if the xl-derivative is kept on one side of equation (5.11)
while the rest of the terms, taken at a mid-point station, is put
on the other side, involving iterations for A2e'

After the plane-of-symmetry solution has been obtained, the cor-


responding boundary-layer parameters are evaluated from equations
(3.1) to (3.15) in slightly reduced form.

The physical shear stress component in the xl-direction reads

(5.13) *13
t

while that one in the x 2 -direction vanishes, since v 2 = O. The


derivative with respect to x 2 , however, exists because of A2:

(5.14 ) *23 r--- 2


(t )'2=-I.L>,a 22 A'3

The heat flux normal to the wall is determined from equation


(3.3) :

(5.15) q
*3

- 39 -
The partial differential equations which have to be solved in
order to compute the displacement thickness discussed in chapter
3, reduce to ordinary differential equations due to the assump-
tion of symmetry.

The three-dimensional mass-flow displacement thickness °1 is


computed from equation (3.5) with:

(5.16)

with

6 1
(5.17) 61 1
x
f
0
( 1 - ~) dx 3
Pe Ve

6 2
(5.18) 61 2 =f ( 1 - ~) dx 3
x 0 Pe Ae

The expression (3.8) for the equivalent inviscid wall-outflow


velocity reduces to:

(5.19 )

The first-order partial differential equations (3.10) and (3.11)


for the momentum-flow displacement thickness °2x a ' a = 1,2 become
ordinary differential equations, too:

- 40 -
(5.20) (v') 2 ( O2 - O2 )]" +
[k O' P e e
,
Pe ve A2e (0 2 , 2 -
x'x'

0
x'

) +
+ k O' 2x'
x x
(v') 2 (6 2
+ k O' k" Pe e
x'x' - 62
x'
) 0

Note that the last term disappears if a" # a" (x,) because then
k" = 0 (see eq. ( 5 . 2) ) .

(5.21) A2
[k O' Pe v e e
, (0 2 - O2 )]" +
x'x 2 x2

Pe (A2) 2 (0 - O2 ) +
+ 2 k O' e 222
,
+ - - k 2 ,'2 P e
x x
(v') 2 (6 2
x2

- 0 ) +
e 2x'
k O' x'x'

k22 Pe ve
, A2 [6
2 - 2"
, (6 2 + O2 ) ] 0
+ k O' e
x'x 2 x' x2

It is interesting to note that equations (5.20) and (5.21) can


be integrated in a decoupled fashion while equations (3.10) and
(3.11) are coupled. The two-dimensional momentum flow displacement
thicknesses are determined by:

o
(5.22) f [1 - a,a = 1,2
o (no summation!),

where F = v' for a,a 1 and F ='A 2 for a,a = 2.The energy-flow
displacement thickness 6 3 , finally, is the solution of the fol-
lowing first-order ordinary differential equation:

(5.23) o ,

with
0 E Fa
(5.24) 63
xa
f
0
(1 - --I
E Fa
a = 1,2
e e (no summation!),
1
again with F = v for a = 1 and F = A2 for a = 2.

- 41 -
The initial data needed to start the integration of these
ordinary differential equations are given by the stagnation-point
values from chapter 4.2.

5.2 Infinite-Swept-Wing Flow

The essential feature of the flow about infinite swept cylinders


and wings is the independence of the flow properties from the
spanwise direction (which is a "natural" coordinate for infinite-
swept-wing flows) because the isobars are parallel to the leading
and trailing edge. Therefore the spanwise derivatives of the flow
(as well as the geometrical) variables are identical zero, and
the spanwise coordinate disappears from the governing boundary-
layer equations resulting into quasi-two-dimensional solutions.
The boundary-layer, however, is truly three-dimensional because
of the sweep of the wing and of the variation of the flow across
the isobars. The assumption of irrotational inviscid flow about
infinite swept wings yields an inviscid spanwise velocity component
which remains constant in the direction normal to the isobars.

From its definition it is obvious that the concept of infinite-


swept-wing flows can only be an approximation to the flow about
finite tapered wings. The approximation improves if the wing re-
sults from a parallel-isobar design because then the isobars are
nearly parallel to the lines of constant chord. In this case the
concept of infinite-swept-wing flow can be applied to the midspan
region of the wing. Close to the root and tip of the wing where
isobars and lines of constant chord deviate markedly from each
other, the concept is not applicable. Nevertheless, unfortunately,
at those locations infinite-swept-wing solutions are often used
to initiate a fully three-dimensional boundary-layer prediction
[36]. Accepting the need for infinite-swept-wing solutions a more
realistic assumption is made with the concept of "locally-infinite-

- 42 -
swept-wing flow" which allows for a chordwise change of the invis-
cid spanwise velocity component at the edge of the boundary layer
(chapter 5.2.3). In fact, the correct external boundary conditions
are used, but the spanwise derivatives in the governing equations
are neglected.

The problem of the infinite-swept-wing boundary layer has been


tackled by many investigators. PRANDTL [71) has looked at the
laminar flow on an infinite swept cylinder. More recently, the
flow about an infinite swept wing has been predicted, for in-
stance, by TREADGOLD and BEASLEY [84), HIRSCHEL [34) (laminar) ,
KRAUSE et al. [50), KRAUSE [46), ADAMS [13), SCHNEIDER [75)
(turbulent), and in particular in the Trondheim Trials [24,27].
HIRSCHEL and JAWTUSCH [37J, for instance, applied the locally-in-
finite-swept-wing approximation to the boundary layer of a finite
swept and tapered wing.

In this chapter, several infinite-swept-wing solutions will be


discussed. Thereafter the initial conditions for the correspondig
quasi-two-dimensional cases are given.

- 43 -
5.2.1 Leading-Edge Oriented Orthogonal Coordinates

x~

Fig. 5.2 Infinite swept wing with surface-oriented orthogonal


coordinate system (x 2 -direction parallel, x 1 _ x 3 plane
normal to leading edge; ~O angle of sweep). Uc (x 1 ) is
the contour angle

The leading-edge oriented orthogonal coordinate system which is


depicted in Fig. 5.2, has the advantage of reducing the governing
equations considerably. Indeed, in first-order boundary-layer
theory no surface-curvature terms are present in the equations,
whose form resembles that for Cartesian coordinates. Due to the
orthogonal coordinates the off-diagonal metric tensor coordinate
a 12 = a 21 vanishes identically. The x 2 -coordinate is rectilinear
while the x 1 -coordinate is curved only within the plane normal
to the leading edge (and x 2 -coordinate). Therefore a 11 and
a 22 = 1 (see chapter 10.1.2), thus the metric factor k01 1 as
well. Most of the rest of the metric factors k mn vanishes except
k16 = k27 = - 1, k41 = k43 = 1.

- 44 -
Introducing these results and neglecting the derivatives with
respect to x 2 in eqs. (2.1) to (2.4) yields the familiar equations
for steady flows (zero-order theory, see chapter 9.1) :

Continuity equation

1 3
(5.25) (p v ), 1 + (p v ), 3 o

Momentum equation for xl-direction

1 1 3 1
(5.26) p [v v ' l + v v'3]

Momentum equation for x 2 -direction

1 2 3 2 2
(5.27) p [v v'l + v v'3] (11 v '3)'3'

Energy equation

(5.28)

+ E ref {v 1 p" + 11
12
[(v '3)
22}
+ (v '3) ]

It is well-known that for laminar incompressible flows the "in-


dependence-principle" holds, i.e. equations (5.25) and (5.26)
are decoupled from eq. (5.27). The boundary conditions at the
edge of the boundary layer are
2
x const, x, ~ 0:

(5.29)
1 v 1 (xl), 2
v v sin !PO const.
e
T T (x'),
e

while the conditions at the wall are usually no-slip conditions


and prescribed temperature or temperature gradient.

- 45 -
The boundary layer parameters in terms of the leading-edge orien-
ted orthogonal coordinate system are easily obtained from equations
(3.1) to (3.15). For the sake of completeness, they are presented
here:
Shear stress in xa-direction, a = 1,2:

(5.30)

Heat flux normal to the surface:

(5.31 ) *3
q - k T'3 .

The partial differential equation for the mass-flow displacement


thickness reduces to that for two-dimensional flow with respect
to v 1 (v~ = 0 at xl = 0) at xl (x 2 = const.):

(5.32)

where

(5.33)

The influence of the velocity component v 2 on the mass-flow dis-


placement thickness 01 is therefore only implicitly existent due
to the fact that the governing equations (5.25) to (5.28) are
coupled. The relation (3.8) for the equivalent inviscid wall-
outflow velocity yields:

(5.34)

The partial differential equations (3.10) and (3.11) for the momen-
tum-flow displacement thicknesses reduce to:

- 46 -
Ii 1 a
f [1 - P v v 1 dx 3 =
e v:
(5.35) a 1,2
o P v~ (no summation!),

because at x 1 = 0 v~ = 0 as well. With the help of the same


argument the relation for the energy-flow displacement thickness
is found from equation (3.14) :

Ii 1
(5.36) f [1 - ~l dx 3
o E
e
v1
e

Note that for a rotational external flow with v e2 = v e2 (x 1 ) it is


necessary to retain the pressure gradient term in the momentum
equation (5.27) in x 2-direction, since otherwise the numerical
solution would not allow to define the free outer edge of the
boundary layer:

1 2 + v3 2 1 2
(5.37) P [v v '1 v '3 - p, 2 + (1.1 v '3)' 3

The pressure gradient term is, in particular, retained for non-


orthogonal coordinates (chapter 5.2.2).

- 47 -
5.2.2 Leading-Edge Oriented Non-Orthogonal Coordinates

X3
1

X7
1

Fig. 5.3 Infinite swept wing with surface-oriented non-ortho-


gonal coordinate system (x 2 -coordinates parallel to
leading edge; x'-x 3 plane is a plane of constant span

,
, , 3'
and parallel to the x -x plane;~O = angle of sweep
of leading edge). a c (x ) is the contour angle.

Consider now the leading-edge-oriented non-orthogonal coordinate


system depicted in Fig. 5.3. The angle ~ between the surface-oriented
coordinates x' and x 2 is the same as that between lines of constant
span and constant chord and therefore equal or less than w/2 on

,
the upper surface for back swept wings and equal or larger than
w/2 on the lower surface. Hence a'2 (= a 2 ,) f 0 forlx I > O. At
the leading edge, not necessarily coincident with the attachment
line, a'2 = a 2 , = O. The metric tensor coordinate a'2 is only a
function of x'. For a planar wing a'2 a2' = const •• The
other metric coordinates a"and a 22 are in first-order boundary-
layer theory equal to one (chapter '0.'.2). Thus the metric fac-
tors k mn , equations (2.'3) to (2.26), are finite except for k'2'

- 48 -
k 13 , k22 and k23 which vanish due to the above assumptions.

Nevertheless, the governing boundary-layer equations (2.1) to


(2.4) here also simplify considerably because of the vanishing
x 2 -derivatives:

Continuity equation

1 3
(5.38) (k 01 P v )'1 + (k 01 p v )'3 o

Momentum equation for xl-direction

1 1
(5.39) p [v v , 1 + v3 v 1 , 3 + k 11 1 2
(v ) I = k 16 1 )
P '.1 + (1.1 v , 3 ' 3

Momentum equation for x 2 -direction

(5.40) P [ V1 2 3 2 1 2 2
v , 1 + v v'3 + k21 (v ) 1= k26 P'l + (1.1 v '3)'3

Energy equation

(5.41 )

+ Eref
{v 1 P'l + 1.1 [k (v 1 '3) 2 + k42 v 1 '3 v 2 '3 +
41

The boundary conditions for equations (5.38) to (5.41) are as


usual. The outer boundary conditions for the velocity components,
however, are different. Either v~ (x') is given or has to be com-
puted from the corresponding velocity component given with respect

- 49 -
to the orthogonal coordinates (chapter 5.2.1 and chapter 10.1.2):

(5.42)

The velocity component in the spanwise direction now varies with


xl :

(5.43) v! (x l )

Concerning the boundary-layer parameters for the non-orthogonal


coordinates shown in Fig. 5.3, the equations (5.30) and (5.31)
for the wall shear stress and the heat flux remain valid. Wall
shear stress and wall heat flux in the present non-orthogonal co-
ordinate system are the same as for the orthogonal system dis-
cussed in the preceding chapter:

(5.44 ) *a3 a = 1,2,


'(

*3
(5.45) q

Relation (5.32) which determines the mass-flow displacement thick-


ness for orthogonal coordinates, has to be replaced by

(5.46)

which still includes the two-dimensional definition for an im-


permeable wall, and equation (5.33) for 01 remains valid. The
xl
equation for the equivalent inviscid wall-outflow velocity reads:

(5.47)

- 50 -
Equations (3.10) and (3.11), which determine the momentum-flow
displacement ticknesses 02 and 02 ' reduce now to:
xl x2

(S.48)

(S.49)

where because of the above assumptions the relation .kll = kOl'l/kOl


has been used to arrive at the form of equation (S.48), and where
k21 = all a 12 '1/a in equation (S.49). The two-dimensional defini-
tions 02x l x a , a = 1,2, are given by equation (3.12), if the non-
orthogonal velocity components v l and v 2 are employed. Note that
equations (S.48) and (S.49) are ordinary differential equations
which are decoupled. With the same argument as for orthogonal
coordinates equation (3.14) for the energy-flow displacement
thickness reduces to the two-dimensional definition:

6 1
(S.SO) f [1 - ~l dx 3
o E
e
v1
e

- 51 -
5.2.3 Locally-Infinite-Swept-Wing Concept

It has been mentioned earlier that the infinite-swept-wing con-


cept can be applied with sufficient accuracy only to those re-
gions of finite wings where the isobars are at least nearly
parallel to the leading edge and to each other. In such a case
the sweep of the equivalent infinite swept wing can be approxi-
mated by the sweep of the leading edge, and the external boundary
conditions can be chosen in the way described in chapters 5.2.1
and 5.2.2. If however, wing planforms (e.g. with a small taper
ratio) result in isobars which diverge or converge markedly, it
is reasonable to drop the infinite-swept-wing assumption in fa-
vour of the locally-infinite-swept-wing concept. This concept
uses only locally the assumption of negligible spanwise flow
variation, and the spanwise velocity component v~ is no longer
related to one sweep angle (chapter 5.2.1, 5.2.2). Instead, the
velocity components v: and v~, as predicted by inviscid-flow
theory or as measured, are used as local outer boundary conditions
[46,13,37]. The locally-infinite-swept-wing concept also allows for
a spanwise variation of the metric coefficients since, in general,
these do not vanish although the spanwise flow variation may be
negligible along lines of constant chord on a tapered wing. The
locally-infinite-swept-wing concept is a good approximation if a
parallel isobar design for a particular case results in nearly
constant pressure along lines of constant percent chord, which
serve as spanwise coordinates. Consider now a station x 2 = const.
of the coordinate system depicted in Figure 10.1. The governing
boundary-layer equations for steady flows then read with a/ax 2 = 0
*a
for all flow variables (note that the concept means v'2 = 0, and
not v~2 - 0, and note also that not necessarily P'2 =0, because
such flow in general would be rotational (see also end of chapter
5.2.1) :

Continuity equation

1 + 2 3
(5.51) (k 01 p v ), 1 + k01 p v + (k 01 p v ), 3 o
where

- 52 -
Momentum equation for xl-direction

3 1 2
(5.52) P Iv 1 v 1 , 1 + V v 1 , 3 + k11
+
(v 1) 2 + k12 v v + k13 (v 2 )2]

1 + 1 2 (V 2 )2]
Pe Iv 1 v e 'l + k11
e
(V~)2 + k12 v e v e + k13 e
+

+ 1
(I! v , 3) , 3

where

+ 1
k12 k12 - 2a
11 all' 2

Momentum equation for x 2 -direction

1 2 3 1 +
(5.53) P Iv V , 1 + V v 2 , 3 + k21 (v 1) 2
+ k22 v v 2 + k23 (v2) 2]

2 1 2 +
Pe Iv 1 v e'1 + k21
e (V~) + k22 ve ve + k23 (v!) 2] +

2
+ (I! v , 3) , 3

where

+ 1
k23 k23 - 2a
22 a 22 '2

Energy equation

(5.54)

{V 1 I (1) 2 1 2
+ E ref P'l + I! k41 v'3 + k42 v '3 v '3 +

where the metric factors k mn are determined by equations (2.13) to


(2.26). The wall boundary conditions for equations (5.51) to
(5.54) are the familiar ones, the external boundary conditions for
the locally-infinite-swept-wing flow read:

- 53 -
(5.55) x, > 0, x
2 t
cons., X3 -> 00:

v'e (x ,), v2 v: (x'), T

2
Note, again, that the relations (5.55) change with x .

The boundary-layer parameters are calculated with equations (3.')


to (3.15):

Shear stresses in x'- and x 2 -direction:

*a3
(5.56) '( =- II la(nn) a = ',2

Heat flux in x 3 -direction:

*3
(5.57) q - k T'3

Mass-flow displacement thickness:

(5.58)

Equivalent inviscid wall-outflow velocity:

(5.59)

where 0, are the well-known definitions of the two-dimensional


xa
mass-flow displacement thicknesses.

- 54 -
Momentum-flow displacement thickness in xl-direction

(5.60)

1 2 1
+ k k p v v [0 - - (0 + O2 ») +
01 12 e e e 2x 1x 2 2 2x 1 x2

and in x 2 -direction

(5.61 )

where

+ kOl +++
k 01 - 2a k 01
11 a 11 '2 1

and where equations (3.12) define the two-dimensional momentum-


flow displacement thicknesses O2
xax~

- 55 -
Energy-flow displacement thickness

(5.62)

with the two-dimensional energy-flow displacement thicknesses


calculated from equation (3.15).

If one chooses another coordinate system, e.g. a locally ortho-


gonal one, but with the x 2 -coordinate still being defined by lines
of constant chord, the governing boundary-layer equations and
parameters are derived from those given in this section, by simply
introducing the correct metric factors.

5.2.4 Initial Data for the Locally-Infinite-Swept-Wing


Solution

The prediction of quasi-two-dimensional boundary layers requires


the specification of initial data at x 1 = 0 (see Figure
5.4). In the case of the plane-of-symmetry flow initial data
are given by the three-dimensional (quasi-one-dimensional)
stagnation point flow solution (see chapter 4). The initial data
for infinite-swept-wing flows are furnished by the corre-
sponding attachment-line flow, a special form of a plane-of-
symmetry flow [8,13,16,37). The attachment line is - as in-
dicated earlier - difficult to define and to specify for the gene-
ral case of a lifting \-ling. HIRSCHEL and JAWTUSCH [37) have lo-
cally replaced the curved attachment line, approximated by the
line where the inviscid chordwise velocity component v~ (Fig. 5.4)
vanishes, in a piecewise manner by lines of constant chord (Fig.
5.4). They have then successfully applied the locally-infinite-
swept-wing concept, including the correspondig attachment-line
solution, to the midspan-region of a finite tapered swept wing
at angles of attack as large as 15 0 in order to predic~ the loca-
tion of the laminar separation bubble and to study the laminar-

- 56 -
turbulent transition location measured in experiment.

In the case of symmetrical flow the non-orthogonal (Fig. 5.3)


coordinate system merges into the orthogonal one (Fig. 5.2) along
the leading edge which then defines the attachment line. Here the
attachment-line boundary-layer equations are given for a non-
orthogonal coordinate system (Fig. 5.4) in the form used by HIRSCHEL
and JAWTUSCH [37]. With the simplifications introduced in chapter
5.2.1, the equations for an orthogonal system can readily be ob-
tained.

L /
1;7 - x'=const.

X
"
Fig. 5.4 Approximation of the attachment line for a lifting
locally-infinite-swept-wing flow at two span stations
(index inf stands for infinite-swept-wing assumption).

- 57 -
The governing boundary-layer equations are obtained from equations
(5.51) to (5.54) by taking the limit for x 1 + 0, with v 1 (x 1 = 0,
x 2 = const., x 3 ) = o.

Continuity equation

(5.63) o.

Momentum equation for x 1-direction after differentiation and intro-


duction of flow symmetry with respect to x 1 :

(5.64)

2
= Pe [(Ae1)2 + k+12 A1e ve2 + k 13'1 (ve )2] + (A1)
~ '3 '3·

Momentum equation for x 2-direction

(5.65)

+ + +
where k 01 ' k12 and k23 are defined in equations (5.51) to (5.53) .

Energy equation

(5.66) cp v 3 T, 3 = pr- 1 + Eref ~ [k 43 (v 2 , 3) 2 ],


P ref (k T'3)'3

where, as before, A1 = V 1 '11


1 . Equations (5.63) to (5.66) have
x =0
to be supplemented with the familiar boundary conditions at the
wall. At the edge of the boundary layer the inviscid-flow predic-
tion yields:

(5.67) O, X2 const., x3 + "':

A1 2 2 2 2 2
e (x ), v ve (x ), T Te (x ).

The boundary-layer parameters become:

- 58 -
Shear stresses in x'- and x 2 -directions:

(5.68) '[
*'3 o ( '[ *'3 ) , , - IJ.
,
~ A '3

(5.69)

Heat flux in x 3 -direction:

*3
(5.70) q =- k T'3

Mass-flow displacement thickness 0,:

(5.71)

Equivalent inviscid wall-outflow velocity:

(5.72)

Momentum-flow displacement thickness in xl-direction (after dif-


ferentiating equation (5.60) with respect to x'):

(5.73)

- 59 -
and in x 2 -direction:

(5.74) kO 1 A 1 2 (0 - O2 ) + k +++
e ve 2x 1x 2 x2 01

These are two algebraic equations which can be solved for and <5 2
x1
O2 after the boundary-layer equations (5.63) to (5.66) have been
x2
integrated. Finally, the energy-flow displacement thickness is
found from the algebraic relation:

(5.75)

For the infinite swept wing the equations for the initial data
follow from the foregoing equations if the derivatives with
respect to x 2 of all metric coefficients or factors are simply set
to zero. This holds for both the orthogonal (chapter 5.2.1) and the
non-orthogonal (chapter 5.2.2) case, for the first, beside this,
a 12 " O.

For the derivation of the above equations it has been assume.d tacit-
ly that k13 = 0 at x 1 = 0, otherwise the first momentum equation
would not have vanished completely. This means that locally the
contour is assumed to be cylindrical, which further implies, that
the attachment line is locally straight and lies on a generator of
that cylinder (see also chapter 12.3). Another point of view would
result if one would deal from the beginning with the momentum equa-
tion for the x '-direction after having it differentiated with respect
, 2 2
to the x -direction. Then one has to assume k13 v v'1 = 0 at
x 1 = O. If k'3 ~ 0, then only v~1 = 0, that is, the flow is symme-
trical with respect to the attachment line. However, this is not
true, in general. In practice therefore an incompatibility arises
between this quasi-one-dimensional attachment-line solution and

- 60 -
the locally-infinite-swept-wing solution continuing it. If the
flow or the surface is not sufficiently symmetrical no continuing
solution can be achieved at all*).

*)
This discussion is due to D. Schwamborn, for details see [78].

- 61 -
6. Initial Conditions for the Prediction of Boundary Layers on
Wings and Fuselages

It is well-known that boundary-layer theory cannot be applied


to the prediction of the complete flew field past an aircraft con-
figuration, because the boundary-layer approach is not valid, for
instance, for the flow along wing-fuselage junctions and along
wing tips (see Fig. 6.1). The boundary-layer on portions of the
aircraft, such as on wings or fuselages, can, however, be predicted
quite well as long as the flow remains attached to the body sur-
face.

It is also well-known that the boundary-layer equations can only


be solved numerically if the employed integration scheme satisfies

Fig. 6.1 Sketch of portions of a complete aircraft configuration


amenable to boundary-layer theory.

- 62 -
limiting wall streamline
direction

Fig. 6.2 Regions of dependence and influence of a velocity


profile within the boundary layer.

the domain-of-dependence condition [41,47,60,72,86,88,89]. This


condition which is particularly important for three dimensional
boundary-layer predictions, is based on the fact that the charac-
teristic directions along which disturbances travel is given not
only by the direction normal to the wall but also by the streamline
directions (see MAGER [60], RAETZ [72], WESSELING [89], WANG [86],
KRAUSE et ale [47,48]). This situation is described in Figures 6.2
and 6.3, the latter being taken from KRAUSE et ale [50]. Fig. 6.2
presents the spatial wedge of influence and dependence of a flow
profile along a wall-normal line. The bounding planes of the up-
stream and downstream wedge regions are wall-normal planes obtained
from considering the characteristic surfaces found from the highest
derivatives of the boundary-layer equations. The opening angle of
the wedge is determined by the sub-characteristics obtained from
consideration of the convective terms [45]. In fact the opening
angle is determined by the external inviscid and the limiting
wall streamlines. The flow profile along the wall normal depends
on the entire upstream wedge region, and influences, on the other
hand, the flow within the entire downstream wedge. This situation

- 63 -
leads to Figure 6.3, an illustration of RAETZ' influence principle
which simply states the following: let initial conditions be speci-
fied on a surface 51 normal to the surface of the body considered,
where the line of intersection does not coincide with the projection
of a streamline. If this surface is of finite extent, then the so-
lution of the boundary-layer equations is completely defined in
terms of those initial conditions only in a domain of influence of
finite extent. This domain is bounded by 5, and two other surfaces
52 and 53 normal to the body, see Fig. 6.3. They are erected over
a limiting wall and/or the projection of an external inviscid
streamline such that the enclosed area of the surface of the body
attains a minimum. The flow outside of this domain of influence
cannot be computed as it is affected from the flow outside of 51'
This is shown in the left part of Fig. 6.3, where the surface
55 intersects the surface 56' and 55 originates outside of sur-
face 5,.

5,

surface
of the body

Fig. 6.3 5ketch of an incremental portion of the region of


influence of the initial data on a surface S of
finite extent: only the solution in part of {hat
portion, say on surface 54' can be determined by
the data on 51 alone.

- 64 -
Since the region of influence is given by the streamline directions,
to be stable, the integration scheme must always be oriented such
that locally the flow is followed, and its numerical domain of
dependence must always include the domain of dependence of the
boundary-layer equations (Courant-Friedrichs-Lewy condition). This
becomes important in three-dimensional flows when the flow lateral
to the main marching direction changes its direction. Such flows
can be handled by the second-order accurate z-shaped integration
scheme of KRAUSE et al. [44,47,481. This scheme is able to march
with conditional stability against cross flow because it uses
downstream as well as upstream information in the cross flow
direction. CEBECI et al. [201 use the characteristic box scheme
which is based on the integration of the boundary-layer equations
in local streamline coordinates within the boundary. layer.
LINDHOUT et al. [581 start their integration, based on some modi-
fied Laasonen [ 91 scheme, from one plane with initial data and
the program computes the flow in the region of determinacy which
is automatically determined when marching downstream. Both latter
methods fail to predict the flow on the outboard portion of the
upper surface of a lifting wing at angle of attack unless some
approximate solution is created along the tip where the flow is
directed inboard. Thus, on wings generally two planes with initial
data are used, along the leading edge and in chordwise direction.

The preceding remarks clearly explain why the boundary-layer pre-


diction for a general wing-body combination has to start at the
stagnation point of the complete configuration. From the stagna-
tion point it has to proceed over the entire surface (Fig. 6.1).
If regions are reached, denoted e.g. by shaded areas in Fig. 6.1,
where boundary-layer theory is no longer valid, the solution can-
not be further advanced. Hence it is necessary, considering the
present state of the art, to treat separately parts of the aircraft
surface. This then poses problems to the investigator with re-
spect to the proper initial conditions to be used. It is not
intended here to discuss the task of patching up the different
flow-field predictions. It is, however, assumed that the in-
viscid flow field, in particular the velocity distribution on the
body surface, is known. Consider first the flow past a body. If

- 65 -
the inviscid flow field is known, the thr.ee-dimensional stagna-
tion point can be determined, and the boundary-layer solution
discussed in chapter 4 can be computed. The construction of that
solution tacitly assumes that at least one symmetry plane exists,
along which a quasi-two-dimensional solutio~ (see chapter 5) is
calculated which can establish initial values for the circum-
ferential (lateral) integration. The integration in the main stream
direction can then be carried out with, for example, the z-shaped
scheme after initial values for that direction have been provided,
for instance near the stagnation point (see below). Often a body-
-oriented coordinate system is used which exhibits a singular be-
haviour at the nose because of the vanishing cross section. This
does not pose any problems for a symmetrical flow because the nose
coincides with the stagnation point (where the computation starts),
but for flows at an angle of attack one has to introduce a special
transformation [17] or for instance an external streamline coordi-
nate system for the nose region in order to circumvent that geo-
metrical singularity (see chapter 10.2).

If a plane-of-symmetry flow doesn't exist on the fuselage, an in-


tegration scheme must be used which needs initial data only on an
upstream surface (e.g., for external streamline coordinates with
an explicit or some fully implicit scheme; see also CEBECI [20],
LINDHOUT et al. [58] or SCHWAMBORN [78]). Initially, such a sur-
face is wrapped around the stagnation point. In order to generate
data on that surface a power series solution [40,80] or the solu-
tion of chapter 4 may be employed with sufficient accuracy if the
considered location is' close enough to the stagnation point. Note
that such a procedure is necessary to generate approximate data
in the neighbourhood of the stagnation point in order to initiate
the integration in the main stream direction also in the case where
a plane-of symmetry flow exists.

Consider now the flow prediction for a finite swept wing, Fig. 6.4.
If the flow field is symmetric on the upper and lower surface
(non-lifting symmetrical wing), the flow along the geometrical
leading edge which then coincides with the attachment or "stagna-
tion" line, is described by the quasi-two-dimensional solution

- 66 -
(see chapter 5.1) denoted in chapter 5.1 with plane-of-symmetry
flow. If, however, the flow is lifting, it can be demonstrated
that there does not exist a quasi-two-dimensional solu-
tion regardless of the definition used for the attachment line
[78] (e.g. locations at which the chordwise velocity component
vanishes or where the flow locally is symmetrical). It is con-
venient then to use the approach discussed in chapter 5.2.4 which
is based on the locally-infinite-swept-wing assumption. An accurate
approach, however, is being made by SCHWAMMBORN [78] along the
leading edge of wings with chordwise symmetry plane (the existence
of the symmetry plane facilitates the generation of initial data),
where the flow in a region encompassing the attachment line is
predicted as follows. The solution is initiated in the symmetry
plane at the stagnation point and progresses in the spanwise di-
rection using the z-shaped finite-difference scheme, at the chord-
wise boundaries the ordinary four-points scheme with correct ad-
vective orientation is employed. Since the scheme requires initial
conditions for the chordwise integration direction for each span-
wise integration step, these are obtained by iteratively solving
the boundary-layer equations with two z-shaped schemes with oppo-
site marching directions for the profiles at two mesh pOints near
the location of the attachment line. Thus, second-order accuracy
is maintained also for the initializing steps in the spanwise di-
rection. Once the flow is determined in the spanwise strip, the
flow in the remaining parts of the lower and upper surface can be
predicted as usual. One of the objectives of the work is to esti-
mate the error when using locally-infinite-swept-wing solutions in
the neighbourhood of the attachment line to start the chordwise in-
tegration. Several results of the investigations by SCHWAMBORN
[78] are discussed in chapter 12.

If a solution is to be found along the leading edge initial condi-


tions have to be provided. For a spanwise symmetrical wing the ini-
tial conditions are given by the stagnation-point solution (chap-
ter 4). For a realistic wing attached to a fuselage (Fig. 6.1) a
stagnation-point solution at the leading edge of the root-section
is an approximation whose quality is difficult to estimate. The
flow field in the neighbourhood of the root section resembles only
to a certain extent that around an oblique cylinder on a flat plate.

- 67 -
x 'teading-Edge" Tip
Solution ~~
I

~
(X)
J?
V
J

"Locally-lnfinite- Swept-Wing"
Solution

x"
''plane-or-Symmetry'' Solution (Root)
Fig. 6.4 Sketch of a wing with an inviscid flow pattern
(planform view) .

In the same sense the assumption of a plane-of-symmetry flow along


the root of the wing is only an approximate one. For highly swept
wings this assumption may introduce numerical difficulties or an
increase in computational effort (too small step-sizes to be taken
in the spanwise direction) for the spanwise integration (36) since
the symmetry assumption leads to a large curvature of the lines of
constant chord in the symmetry plane. Therefore it seems to be more
reasonable to use the locally-infinite-swept-wing assumption along a
line of constant span, discussed in chapters 5.2.3 and 5.2.4 in
order to provide initial data for the spanwise integration (Fig.
6.4) if necessary. If the span station is close enough to the root
section inviscid-flow predictions have shown that the spanwise
velocity component is small enough to restrict the domain of in-
fluence of the wrong initial data to a sufficiently small stream-
wise region. The parabolic character (in span as well as chord-

- 68 -
wise direction) of the boundary-layer equations reduces the in-
fluence of the initial data chosen, anyway.

In the case of flow at an angle of attack the tip region of the


wing will create a vortex with the air flowing inboard on the upper
surface. In order to satisfy the domain-of-dependence condition
there, initial data have to be specified along the tip, and the
spanwise integration has to proceed, in general, in the inboard
direction [18]. In such cases locally-infinite-swept-wing solu-
tions can be used. This approximation influences near the tip
a larger downstream region than near the root section because
the spanwise velocity component is larger. Due to its ability to
march against cross flow the use of the z-shaped difference scheme
doesn't require the locally-infinite-swept-wing assumption close
to the tip if an appropriate solution is known along the wing root.
For the last spanwise integration step, however, another procedure,
for example extrapolation, is needed [42] for negative cross flow
in order to prevent the domain of integration from shrinking in
the spanwise direction while marching in the chordwise direction.
For positive spanwise flow at the tip simply the ordinary four-
point scheme [48,49] is employed [78].

- 69 -
7. Higher-Order Boundary-Layer Equations

If locally the boundary-layer thickness is not small compared with


the smallest radius of curvature of the surface, the pressure gra-
dient normal to the surface - within the boundary layer - may no
longer be small of higher order, and hence no longer be negligible.
This is the best known higher-order effect. There are other higher-
-order effects, see e.g. van DYKE [61, but here only curvature
effects are being considered and described by the addition of terms
to the first-order boundary-layer equations (see chapter 2). The
equations presented here have been derived by ROBERT [31, who
evaluated separately the influence of wall curvature itself and the
wall-curvature gradient, an approach which was found to be necessary
by KUX [11. No attempt is being made to review the numerous sets
of governing equations (including the modeling of turbulence), which
appeared in literature, and the corresponding methods to solve them.

Let K1 and K2 (see appendix A.8) denote the principal curvatures of


the surface under consideration then second-order theory applies
if (see also chapter 9.1)

(7.1a) o (1) < 0 (e) < 0 (t)


Two cases can be distinguished according to the influence of the
wall-curvature gradient [ 31:

I) 0 (1) < 0 (D) < 0 (2)


6
(7.1b)
II) 0 (D) 0 (1)
6

where e denotes the largest of the absolute values of the principal


curvatures:

(7.2)

D represents curvature-gradient terms, and 6 is the boundary-layer


thickness. In the following it is assumed that all metric coeffi-
a
cients are 0(1), whereas all curvature terms, involving b S or b 6'
a Y
are Ole), and all gradient-of-curvature terms, involving bS'n or
b~, , are O(D).
yo II

- 70 -
In order to derive his equations, ROBERT [ 3] introduces shifters
- a concept taken from shell-elasticity theory -, which express
metric quantities associated with the space above the surface in
terms of the surface metric in a very elegant manner. Shifters can
also be understood as relations which transform velocity coordi-
nates for xa-coordinates away from the surface into those given by
the directions of the coordinates on the surface (see below and
appendix A). Here, if convenient, shifters are avoided in the for-
mulation of the governing equations in order to be consistent with
the remainder of the book. Then the equations are given with the
metric factors based on the appropriate a- or g-metric. Note that
the metric factors which contain the a-metric (case I, eq. (7.1b»
are denoted by k mn = k (xl, x 2 ), and those which contain the
mn _ 1 2 3
g-metric (case II, eq.(7.1b», byk =k (x, x, x). The ve-
mn mn
locity coordinates are in either case given without overbar:
*a a 1/2
v v /(g(aa» • It is worth to note that curvature-gradient
terms are usually not regarded in second-order theory.

ROBERT's equations for general three-dimensional flow in the pre-


sent notation read - without boundary - layer stretching:
Continuity equation for case I

(7.3) 1 2 3 o ,
kOl P't + (k 01 P v ), 1 + (k 01 P v ), 2 + (k 01 p v ), 3

and for case II

( 7 • 4) 1 2 3
kOl P't + (k 01 P v ), 1 + (k 01 p v ), 2 + (k 01 p v ), 3 +

-+ 1 -++
+ k01 P v + k01 P v 2 0

where

(7.5)

Momentum equations for xa-direction (a = 1,2), case I

(7.6) a v1 a + v2 a + v3 a
P V't + v'1 v'2 v'3 +

- 71 -
-+ -+
(V 1 )2 + ka2 1 2 (v 2 ) 2 1
+ ka1 v v + k+
a3

a
ka6 P'1 + ka7 P'2 + ~ (\I v'3)'3
ref

where

-+
(7.7) ka1 = Aa1; 1
1;
MO { 0 1 }

-+
ka2 Aa MO { 0 1; 1 } + Aa MO { 0 l; 2}
l; 2 1; 1

-+
ka3 = Aa1;; MO
2 fo
l;
2}

For case I I one obtains

a a 2 a 3 a
(7.8) p [ V't + v 1 v'1 + v v'2 + v v'3 +

+ k (v 1 )2 + ka2 v 1 v 2 + ka3 (v 2 )2
a1

1
ka6 P'1 + ka7 P'2 + Re ref (\I v~ 3) , 3

Momentum equation for x 3 -direction for cases I and I I

(7.9) p [ k31 (v 1 )2 + k32 v 1 v 2 + k33 (v 2 ) 2 1

Energy equation for cases I and II

(7.10)

1 1
Pr ref Re ref

- 72 -
+ __11_
Re ref

The shifters of the first Milv and second A6E kind are defined in
appendix A.14, see also below. Note that the velocity coordinates
appearing in the above equations are those measured with respect
to the local xi-coordinates away from the surface.

Wall shear stresses and heat flux are defined as for first-order
boundary layers, see chapter 3. The situation is different for in-
tegral parameters because of the skewing of the surface-parallel
coordinates away from the surface, see appendix A. If the skewing
is expressed by the variation of the angle ~ between surface-paral-
lel coordinates along a surface normal:

(7.11) cos ~

it is clearly seen from eq. (7.11), that ~ is constant in general


1 2
for x , x = const. only if the curvature vanishes identically.
Due to this skewing the velocity coordinates va (xl, x 2 , x 3 ), which
are measured along the general base vectors Sa (xl, x 2 , x 3 ), a =
= 1,2, are usually different from those associated with the surface
base vectors a (xl, x 2 , 0) which point in the directions of the
-a
coordinates in the surface. It is, however, feasible to transform
tensor coordinates with respect to ~a into those with respect to
~a by applying the transformation laws displayed in appendix A.
Such a transformation is necessary to integrate vector quantities
across the boundary layer as is required for integral parameters.

Consider different representations of the velocity vector which is


split into surface-parallel (a) and - normal (3) components:

1 2 3 1 2 3)
(7.12) Y(a) (x,x,x ) = v a (x,x,x Sa (x 1 ,x 2 ,x 3 ) =

- 73 -
1 2 3 1 2 3 (1
V~ (y) x ,x 2 ,x 3)
~y
v (x,x,x) a (x,x,x 0)
-y

i'
v ~i'

~ (3)

Here and in the following the circumflex accent ~ indicates the


tensor coordinates required for the determination of the integral
parameters, and i' means the Cartesian coordinates of the sur-
face-parallel vector v(a). Two ways are possible to arrive at ~y:
- ai' ~y
one way is to transform v into v and then into v by making
use of the transformation relations for the surface-oriented coordi-
3
nates for x = O. The other possibility, chosen here, is to express
a
Oia in terms of a_y • This is easily done by means of equation (A.16),
which is written here as follows:

(7.13)

with the shifter of the first kind [ 31:

(7.14)

From eq. (7.12) the velocity coordinates needed then follow:

(7.15)

~* 1/2 ~ y
Note that v Y = (a()) v Hence, once the velocity coordinates
a YY . ~Y
v are determined, the velocity coordinates v , based on a coordi-
1 2
nate system which is constant along a surface normal x , x =
= const., can be obtained as is required for the integration of in-
tegral parameters.

KUX [11 arrives by means of flux considerations at the formula-


tion of the viscous displacement problem (see also chapter 3):

f (t'" e
05
Aa ". Aa 3
(7.16) ve - tV) M dx - o
o
- 74 -
". Aa. .
Here t v ~s the boundary-layer flow quantity displacing the ex-
A An
ternal flow quantity t v (note that only the tangential fluxes
e e A A3
are being considered, the fluxes t v are still small of higher
A An
order). The quantity t v may be, see chapter 3, the density flux
(t = p), yielding the mass-flow displacement thickness 01 from eq.
A

(7.16), or a momentum flux (t


= p ~a), yielding the momentum-flow
displacement thickness 02 a ' or an energy flux (t = E), yielding
A

the energy-flow displacem~nt thickness 03' The initially unknown


displacement thickness is denoted by Qt' Note that the external
A An 3
flow quantity te ve is a function of x in higher-order boundary-
-layer theory. The quantity M in eq. (7.16) is the determinant of
the shifters, see also Appendix A.14:

(7.17) M

where the Ka are the principal curvatures (see Appendix A) .


Moreover, M relates the determinant of the g-metric tensor to that
of the a-metric tensor 71 :

(7.18) M = (g/a) 1/2 0)

Eq. (7.16) is an integro-differential equation for the unknown dis-


placement thickness Qt' All flow variables, including the shifted
ones, are known or can be determined, either from experiment or
from a prediction. The determinant of the shifters is a function
of the surface properties and of x 3 only. Due to the integration
with respect to the wall-normal direction the partial differential
equation (7.16) becomes an equation which is to be integrated on
the surface. Hence the rules of differentiation for surface tensors
apply [ 2, 71, e.g.:

(7.19)

- 75 -
Introducing eqs. (7.19) into (7.16) yields the definition equations
for the displacement thicknesses:
Mass-flow displacement thickness
61
Ct. 1 A1
J
_ 3
(7.20) [ k01 1 Pe ve M dx ) I ,1 +
0
61
A2
+ [ k01 (d 2
1
_
f
0
Pe v e M dx 3 ) I ,2 0

where

(7.21) n = 1,2

62 1

+ [ k01
(d 12
2
_ Jx Pe ve
A1 A2
v e M dx 3 ) I , 2 +
0

62 1

+ k01 k11 [ d 11
2
- J x Pe (~1)
e
2 M dx 3 1 +
0

62 1
1 A1 A2
+ k01 k12 d 12 - '!
2
J x Pe v e v e M dx 3 +
0

'"1 '" 2 3
+ P eVe ve M dx ) 1 +

- 76 -
62 2

+ k01 k13 [
f:,22 _
2
J x Pe (~:)2 M dx3 ] 0
0

where
<5
Aa A6 Aa AB
(7.23) f:,aB = (Pe v e ve - P v v ) M dx
3
a,B 1 ,2
2 J
0

Momentum-flow displacement thickness for x 2 -direction

A1 A2 ·3
(7.24) PeVe ve M dx ) ] , 1 +

62 2
+ [k 01 (f:,;2 - Jx Pe (~:)2 M dx 3 ) )'2 +
o

62 1
+k01k21 [f:,~1_ JX Pe(~~)2Mdx3] +
o

+ +

Energy-flow displacement thickness

- 77 -
(6 1 _
°3 E Al
ve M dx 3 ) 1 , 1 +
(7.25) [ kOl 3 f
0

°3 E vA2 M dx 3 )
[ kOl (6 23 -
+
0
f e 1 ,2 0

where
cS
Aa Aa
6 a3 ) M dx 3
(7.26)
f
0
(Ee ve - E v a 1 ,2

The solution of the above equation would proceed in an iterative


manner: integrate the equations in the surface coordinates e.g. in
finite-difference form, and change locally cSt until the dif-
ferential equation is fulfilled. The upper integration bound cS does
not pose a problem here, but may be defined as discussed in chapter
8, if applicable. The equivalent inviscid source distribution and
surface-normal mass can be treated in analogy to first-order theory
(chapter 3). KUX [ 11 introduces reference quantities of the form

(7.27) it 1 ,2,
o

such that the equations are first solved for the quantity 6t • How-
ever, in contrast to the situation in first-order boundary-layer
theory, it is not evident for quantities like the momentum flux
and even the mass flux, how the reference quantities have to be
chosen such that a unique value for cSt results. Instead it is sus-
pected that different values cSt will result according to the value
of a in eq. (7.27).

Note that the first-order-theory relations of chapter 3 are found


from the above equations by neglecting the dependence upon x 3 for
the external inviscid flow and for M:

- 78 -
(7.28) a 1,2 M -

There is still the question how to find the external inviscid flow
field v: (x 1 , x 2 , x 3 ), if the inviscid flow is not provided numeri-
cally by means of a field method but if only v: (x 3 = 0) is given.
In potential flow the condition of irrotationality can be used to
construct locally the necessary data by taking into account the
surface curvature of the body (see e.g. [22,29]). For supersonic
external rotational flow TASSA, RESHOTKO and ANDERSON [83] use a
coupling procedure with the field solution suggested originally by
FERRI and DASH.

- 79 -
8. Thin-Layer Approximations of the Navier-Stokes Equations

8.1 Introduction

The governing equations for time-dependent three-dimensional laminar


compressible non-reacting flows in terms of general curvilinear co-
ordinates read in tensor notation [ 21 with all indices = 1,2,3:
Continuity equation

(8. 1 ) o

where the symbol { } denotes the Christoffel symbol of the second


kind [ 7,121, and the bar indicates that the term is used at a
general field point (~ee eq.(A.63) on page 215).
Navier-Stokes equations for xi-directions

(8.2)

Energy equation (in terms of the temperature T for a thermally per-


f ec t *)
gas )

(8.3) p c p (T , t + vi T , i )

where the stress tensor T ij is given for Newtonian fluids as follows:

l' i
(8.4) g J [( -p + A 0) <5 1 +

*) Note, that in equation (4.56) of [21 ~'ij (~ = T) must be re-


placed by ~Iij since grad ~ gij (~'i)lj = gij (~'ij - ~'l {i\})·
The same has to be done in 31 .

- 80 -
The covariant derivative of the stress tensor is

(8.5) I], = g ] (-p


i'
+ ). 0)" + gm~
]'"
, , - ,
[II (v],
m
+ v n { ] }) 1
mn 'j
+

+ glj [IL (v~l + v n { i })I +


n 1 ' j

+ II { gmi [( v P, + v t { P }) { j }
... m m t j p -

The dissipation function ~ in the energy equation is defined as


follows:

(8.6)

In equations (8.5) and (8.6) the symbol). denotes the second vis-
cosity coefficient which is related to the bulk viscosity IL' in the
following way [ 31

2
(8.7) IL' - '3 IL

- 81 -
and the symbol 0 is

(8.8) o v~m + vk { m }
m k

Note that all (latin) indices in equations (8.1) to (8.8) run from
one to three, when the equations are written out explicitly.

In terms of the locally monoclinic surface-oriented coordinates


(see appendix A.2), which are generally being used throughout this
book, equations (8.1) to (8.8) become the following. As usual the
indices for the two surface parameters xl and x 2 and the associated
quantities are greek letters, running from one to two, while the
index for the surface-normal coordinate x 3 is given explicitly 3).
a *a 1/2
The velocity coordinates are defined as v = v /(g(aa))
Continuity equation

(8 . 9) a
P't+(pv)'a+ a } +p v 3 { a3
Pv y { ay a } +(pv)'3
3 o

Navier-Stokes equations for xa-direction, a 1,2

Navier-Stokes equation for surface-normal x 3 -direction

P
T3j I.J

Energy equation

( 8 • 1 2) (T 't + Va T 3 )
P cp 'a + v T'3

- 82 -
+ gaa k T, Q
'a IJ

The stress tensor reads:

Taj ga j 3
(8.13) (- p + A E/) + fJ. [gay (6 j v n'y + .sj V'y) +
n 3

+ gC j a g 3'J v,a 3 + g 6'J gay n 3


v'c + (g.sY'n v + g6y'3 v )1

and

(8.14 ) T3 j g 3'J (- P + A e+) +

3 + gC j
+ fJ. [.s Jn' vn'3 +.sj3 v'3 3 3j 3 1
v'c + g v'3

The covariant derivative of the stress tensor becomes:

+
(8.15) Tllj gaa (- p + A e ), a +
j

- 83 -
({-e-} { B} {-3-} { B })
£Y Be + £Y 83 -

a {-3-} _ v Y
- v/3 I) B

and

(8.16) T3j I.
]

£
+ v

- 84 -
3 3
- v 'y - v, 3

The dissipation function ~+ reads:

(8.17) ~+

+ gya [vB a + ve: vn {-B-} {-o-}


goB 'a V'y a e: y n +

+ v3 vn n/ 3 } {yOn} + {-B }
a n {y03})
+

(v3) 2 {aa } 3
+ 3 {/3} + 2 v n va'a {yOn} + 2 v va'a {/ 3}] +

3 3 e: -3- 3
[v'a V'y + v v n {a e: l { / n} + 2 v n v 'a. { y 3 n l] +
+ gya

+
B a e:
goa [v, 3 v, 3 + v v n {/J a
{/ n l + 2 v n v, 3 {3 On l] +

-3-
+ v y v a [{ a \} { a } + {a \} {/ a} + {3 \} {a ol] +
a a

- 85 -
and e+ is given by

(8.18) e+ Y 3
v'Y + v'3 + v
O{YI
Y of + v
3{Yl
Y 3J

Due to the splitting of the latin indices into greek ones and 3,
equations (8.9) to (8.18) appear to be more voluminous than the ori-
ginal equations, but, in fact, the number of terms is considerably
reduced due to the reduced metric tensor (A.28). This is reflected
in table 8.1. Table 8.1 compares the number of terms in the gover-
ning equations for general curvilinear coordinates with those for
locally monoclinic surface-oriented coordinates, and with those
for Cartesian coordinates. These numbers result if all differen-
tiations are carried out, and if only algebraic factors are taken
as one term. Furthermore, Christoffel symbols are counted as one
quantity. From the numbers, displayed in table 8.1, it becomes
clear that a solution of the Navier-Stokes equations (including
continuity- and energy equations) is evidently impracticable for
general curvilinear or even locally monoclinic surface-oriented co-
ordinates.

There are several approaches which considerably reduce the compu-


tational efforts needed for the integration of these equations. One
is to use orthogonal curvilinear coordinates which, however, is not
always possible in three dimensions. Most of the approaches profit
from the prominent physical feature of viscous flows, namely that
mostly viscous effects are important only in a region close to the
surface of a configuration, in the boundary layer, and that the
boundary layer is thin at high free-stream Reynolds numbers. For
such flows it is reasonable to retain only the boundary-layer terms
in the governing equations for external as well as for internal
flows. In many cases results of computations with such simplifi-
cations agree very well with experimental data for laminar boun-
dary-layer flows even if separation is present. This is expected
to hold also for flows with massive separation associated with vor-
tices, since viscous effects are, in general, small once separating
layers have left the surface and have formed vortices. The decision
to consider boundary-layer terms only is supported in the case of

- 86 -
Table 8.1: Cacparison of the number of tenns in the governing equations with respect to different coordinate systems

advective **) **)


tenus pressure viscosity bulk viscosity heat conduction sum

G L-M C G L-M C G L-M. C G L-M C G L - M C G L - M C


nates
equations
~
continuity 16 13 7 - - - - - - - - - - - - 16 13 7

Navier-Stokes ~3 12 4 3 2 1 729 354 14 99 48 6 - - - 844 416 25


1 2
x,x-
m:::mentum
co
-.J

Navier-Stokes 3 8 4 3 1 1 729 164 14 99 24 6 - - - 844 197 25


x 3-m:::mentum

energy 4 4 4 4 4 4 1413 386 27 117 51 9 45 22 6 1583 467 50


equation

stress tensor - - - 9 5 3 837 81 30 108 45 9 - - - 954 131 42

*) G: general, L - M: locally m::moclinic surface-oriented, c: cartesian coordinates.

**) the tiIre-dependent tenn is included.


high-Reynolds number turbulent flows by the state of the art in
turbulence modeling for practical applications. For the near fu-
ture at least it seems to be impossible to model much more than
the Reynolds stresses derived in boundary-layer theory . Note that
in classical turbulent boundary-layer theory the viscous (molecular)
terms have been neglected since the Reynolds stresses are several
orders of magnitude larger than the molecular shear stresses. This
is appropriate as long as the resolution of the flow field closest
to the surface of the considered body is coarse, and if one deals
with fully turbulent flows only. A complete flow-field description,
however, includes laminar, transitional and turbulent flows, so
that it is, in general, necessary to retain the molecular shear
stress terms. To date, it is not yet clear how to simulate properly
transitional flows.

Because the boundary-layer terms dominate the viscous effects near


the surface, they will probably take care also of minor viscous
effects away from the surface although other terms might become im-
portant (see chapter 8.2). The importance of the different terms
in the governing equations at moderate or small Reynolds number
still remains to be researched. A careful evaluation of the terms
in the governing equations will still remain necessary even when
the trend to approximate turbulence by means of the large-eddy si-
mulation [73,111] can be realized for practical purposes with the
advent of computers much larger than projected today.

Once it is recognized that the (surface-normal) boundary-layer terms


are generally the most important terms for laminar and turbulent
flows, it is evident that a locally-monoclinic surface-oriented co-
ordinate system is most suitable for the prediction of viscous
flows since those terms are then easily identified. Unfortunately,
the establishment of a surface-oriented general coordinate system
in three dimensions is a very difficult task, see for example [104,
106] •

Now a brief and certainly incomplete review of attempts to reduce


the computational effort needed for the integration of the full
governing equations for three-dimensional viscous flows, either

- 88 -
laminar or turbulent (then solving the time-averaged form of the
equations) is given. Most investigations aim at predicting steady-
-state flows although, recently, it became feasible to integrate
simplified forms of the time-dependent equations at reasonable cost
for research purposes [111,114,116,125,126). The methods cited are
based on time-accurate finite-difference schemes, and therefore
have the potential to compute unsteady flows for appropriate initial
and boundary conditions. In general, however, these methods are used
to obtain the steady-state solution for large times in an asympto-
tic manner (transient unsteady approach) because large time-steps
may then be taken since the evolution of the solution with respect
to time is of no interest unless the rate of convergence is affected.
In most cases the transient unsteady approach is still too time-
consuming to be attractive for practical purposes, and the same
holds for the alternatives for steady-state solutions, namely global
iterative methods. Therefore, to date approximate equations, which
are loosely called "parabolized" Navier-Stokes equations, are pre-
ferably employed since these result in a marching procedure for the
integration. Note, that the notion "parabolized" does not necessari-
ly mean that mathematically the equations are of parabolic type.
The marching direction must coincide with that streamwise direction
where separation is absent, and the viscous diffusion with respect
to that direction must be negligible in comparison with those in
the transverse directions. The loss of ellipticity in the stream-
wise direction can, however, lead to stability problems if the
streamwise pressure gradient is included in the solution algorithm
as is discussed in [109,128,131). The work by WALITT and TRULIO
(133), which is one of the earlier developments, is similar to the
space-marching methods, but requires a further restriction namely
that the body be slender and that the velocity component in the
direction of the axis of the body therefore remains approximately
equal to its free-stream value. Using HAYES' equivalence principle
they transform the steady three-dimensional problem into a time-
dependent two-dimensional problem where time replaces the axial
marching direction. Applications of the usually employed marching
techniques, which differ from each other essentially in the way
the pressure gradient in the marching direction is being treated,
are given, for example, by [107,109,110,113,119,120,121,124,128,

- 89 -
127,131].

The computational effort required for solving the viscous-flow


equations can be reduced considerably if the viscous three-dimen-
sional flow in question is quasi-twa-dimensional (see chapter 5).
One example is the supersonic flow past pointed cones at angle of
attack, which can be tackled by means of the conical-flow assump-
tion [122]. Another example concerns the assumption of azimuthal-
invariant flows on body geometries of axisymmetric type or, in
particular, on infinite swept wings [123]. It should be noted that
in all references cited the transient unsteady approach is used.

The transient unsteady approach is, as was mentioned earlier,


generally used for the integration of the complete three-dimensio-
nal viscous-flow equations. The equations are, however, simplified
insofar as the specific body geometry is introduced explicitly,
see e.g. [114,115,116]. In the cases where general transformations
are incorporated into the equations the "thin-layer approximation"
is preferred. The "thin-layer approximation", a notion which was
coined by BALDWIN and LOMAX [108], reduces the Navier-Stokes equa-
tions to the boundary-layer equations with the pressure as dependent
variable, and supplemented with the momentum equation in the surface-
normal direction such that the complete flow field can be predicted
without the need for an inviscid-viscous coupling procedure. The
justification for the thin-layer approximation is that many viscous
flows have one dominant direction, the wall-normal direction, con-
cerning the importance of flow gradients, which means that the com-
putational mesh should be locally surface-normal in the physical
space. A further, more restrictive justification is that present-
day computers allow to resolve the flow field in only one direction
in sufficient detail. The applicability of the concept in three-
dimensional external flow about convex bodies is demonstrated by
[116,117,125,126]. For the case of flow past an inclined body of
revolution HUNG [116] reports that the solution of the complete
Navier-Stokes equations in cylindrical coordinates requires 25 per-
cent more time than the version with the thin-layer approximation
incorporated, while the results agree within 3 percent. To the
authors' knowledge the thin-layer approximation has only been applied

- 90 -
in context with the transient unsteady approach. The form of the
steady-state thin-layer Navier-Stokes equations resembles higher-
order boundary-layer equations; the boundary conditions far away
from the surface are, however, quite different, since there the
thin-layer Navier-Stokes equations reduce essentially to the Euler
equations, and require appropriate boundary conditions.

8.2 Thin-Layer Approximations

8.2.1 Boundary-Layer Like Flows

Consider a flow field at a smooth body where a boundary layer de-


velops to which the classical boundary-layer assumption applies in
major portions of the flow but which may separate. Then in the
thin-layer approximation of the Navier-Stokes equations all viscous
terms except for the classical three-dimensional boundary-layer
terms are neglected; In the case of turbulent flows, the approxi-
mation of the Reynolds stresses has, of course, to be incorporated
in the time- or mass-averaged governing equations. There are loosely
two phenomenological points of view by which one can look at the
thin-layer Navier-Stokes equations. One is to take them as a set of
Euler equations supplemented by boundary-layer viscous diffusion
terms, the other is to look at them as a set of equations which re-
semble boundary-layer equations (with the pressure as unknown va-
riable) supplemented with the wall-normal momentum equation. Note
that the thin-layer version as used by investigators at NASA-Ames
(e.g. [118,125,126)) differs from the following version with re-
spect to its form. The Ames-version keeps the original dependent
variables, typically Cartesian, and transforms the independent
variables only, thus introducing contravariant velocity coordinates
which exist together with the Cartesian ones. The present version
exhibits, consistent with the remainder of the book, a contravariant
form. The advantage of the mixed form is that only first deriva-
tives of the transformation relations are required while the contra-
variant form needs second derivatives, and thus a smoother approxi-
mation of the metric properties of the surface.

Based on equations (8.9) to (8.12) for locally monoclinic surface

- 91 -
oriented coordinates, the following thin-layer Navier-Stokes equa-
tions result. Note, that here the boundary-layer stretching is being
omitted, and that the pressure p is being referenced with a refer-
-
ence pressure Pref' a scaling which is usually used for computa-
tions of compressible flows. Furthermore, the metric factors k ,
mn
which are the Christoffel symbols of the second kind, contain all
metric information required for the problem in consideration. The
bar over the metric factors indicates the use of the g-metric that
- - 1 2 3 a *a 1/2 3 *3
is k mn k mn (x , x , x ), and v = v I(g(aa» , v = v
Continuity equation

1 2 3
(8.19) kOl P't + (k 01 P v ), 1 + (k 01 P v ), 2 + (k 01 P v ), 3 O.

a
Navier-Stokes equations for x -direction, a = 1 ,2

a 1 a 2 a 3 a
(8.20) P [v't + v v'l + v v'2 + v v, 3 +

1 2
+ k
al
(v 1) 2 + k
a2
v v + k a3 (v 2 )2 +

1 3 2
+ k v v + k v v3)
a4 as

1 a
(k a6 P'l + k + (I! v, 3) , 3
)2 a7 p, 2) ~
Yref (M ref ref

Navier-Stokes equation for x 3 -direction

3 1 3 2 3 3 3
(8.21) P [v't + v v'l + v v, 2 + v v'3 +

(v 1 )2 + 1 2 (v2) 2)
+ k31 k32 v v + k33

1 3
2 P'3 + ~ (I! v, 3) , 3
(M ref ) ref
Y ref

Energy equation (thermally perfect gas)

1 v2 3)
(8.22) c p P [ T't + v T'l + T'2 + v T'3

- 92 -
+
Pr ~ (kT'3)'3
ref ref

1 2 3
P't + v P'1 + v p, 2 + v P'3
+ E +
ref

+ --~-
1 2 1 2 2 2
Re ref Ik 41 (v'3) + k42 v, 3 v, 3 + k43 (v'3) 1

These equations have, of course, to be supplemented with molecular


or turbulent transport relations, the equation of state, and ini-
tial and boundary conditions, see also chapter 2. In eq. (8.21) one
viscous term is arbitrarily retained, although it is small of
higher order compared to the viscous te~m in eq. (8.20.). This is
done for numerical reasons, but depending on the scheme used, it
might not be necessary. The reader may verify that there are more
terms in expression (8.16), some multiplied with the bulk viscosity,
which are of the same order of magnitude as the one term retained.
Depending on the problem another choice might be more appropriate.

The additional metric factors kmn are defined as follows (see also
append ix A. 1 3) :

(8.23)

(8.24)

(8.25 ) k24 g (g 11 g12'3 - g12 g11'3)

1
(8.26) k25 g (g 11 g22'3 - g12 g12'3)

1
(8.27) k31 - "2 g11'3

(8.28) k32 - g12'3

1
(8.29) k33 - "2 g22'3

In cases where the domain of integration of the Navier-Stokes

- 93 -
equations is restricted essentially to a very thin boundary-layer
sheet, e.g. in a combined Euler-Navier-Stokes computation, it might
be sufficient to use the a-metric, that is to replace the general
metric factors k- mn by k mn = k mn (x 1 , x 2 ) which depend upon the
metric tensor of the surface only. If,however, a global application
of the Navier-Stokes equations is intended, the g-metric must be
retained in general form (k mn ).

8.2.2 Corner-Layer Like Flows

Corner flows are characterized by two directions in which viscous


diffusion is equally important, while viscous effects in the third,
streamwise direction are negligible. Due to the limitation of lo-
cally monoclinic surface-oriented coordinates with respect to flows
along concave surfaces, practical concave corner flows cannot be
handled. Only if the corner has a sufficiently large radius of cur-
vature, like in Fig. 8.1, such surface-oriented coordinates can be
used. For the concave corner, see Fig. 8.1, case a), a Cartesian

@
const.
X 1 X 2 :=
,
-+--+--1---4 x '' x 3=const.

Fig. 8.1: Smoothed out corners, a) concave corner with matching


1 1I
Cartesian coordinates, bl cQnvex corner, x , x =constant.

or other suitable coordinate system has to be matched in order to

- 94 -
prevent the locally monoclinic coordinates from converging and
intersecting (see also chapter 1). For the case of a convex corner,
Fig. 8.1, case b), the surface-normal coordinates diverge away from
the surface, and there may be flow cases which require an additional
outer coordinate system in order to provide the resolution necessary
for the flow field.

Since viscous diffusion has to be considered in both lateral direc-


tions for corner flow regions the thin-layer approximation of the
Navier-Stokes equations for boundary-layer sheets has to be supple-
mented by additional terms. In eq. (8.20) for the x 1-direction
(a = 1) the one viscous term has to be replaced by two terms:

(8.30) 1 1 1
-
Re- - [(Il v'2)'2 + (Il v'3)'3]
ref

where it is still assumed that x 1 points in the direction where


viscous diffusion is negligible. In eq. (8.20) for the x 2-direc-
tion (a = 2) the viscous term becomes

1 2
(8.311 - v'2)'2
Re- - (Il
ref

For eq. (8.21) the above remark also holds so that here the equa-
tion can remain unchanged. In the energy equation (8.i2) not only
the viscous, but also the heat-conduction terms are different now.
The viscous terms become

(8.32)

and the heat conduction is described by

1
(8.33)
Pr ref Re ref

Numerical predictions for corner flow regions are reported by,


for example, [114,115,120,129] and by most of the investigators
of internal flows in rectangular ducts, see e.g. [109,124].

- 95 -
Note again, that the form of the equations given in this section
is somewhat arbitrary and represents suggestions, since the equa-
tions are derived heuristically. Depending on the flow field under
consideration, it might become necessary to include additional terms
to improve the prediction. It is also.noted that neither the poten-
tial nor the shortcomings of thin-layer approximations are thorough-
ly researched. Investigations for two and three dimensions, conduc-
ted at NASA-Ames Research Center, are, however, very encouraging
concerning the thin-layer approximation discussed in chapter 8.2.1
[116,117,118,125,126,130,131].

For general cases, which occur if the viscous effects in all three
spatial directions are of the same order of magnitude, the above
ideas can be extended further. In general such a situation will
exist only locally, e.g. at the leading edge-tip junction of a wing,
at a wing-fuselage junction or near highly skewed separation lines.
Suppose that the local characteristic lengths in all directions are
of order E, with E < 1, and that all metric properties as well as
the velocity components are of order 0 (1). Then all second-order
viscous terms are of order 0 (1) larger than first-order terms, and
E
of order 0 [(1)2] larger than second-order terms. This situation
E
can be used to derive an appropriate approximation of the Navier-
Stokes equations.

8.3 Integral Properties

For two-dimensional flows the boundary-layer thickness and integral


quantities in form of displacement or loss thicknesses are used
in correlations and, in particular, in turbulence models. The
definition of the integral quantities comes from boundary-layer
theory where the location of the boundary-layer edge is easily iden-
tified because the flow field is split into an inviscid and a vis-
cous portion. Then the boundary-layer edge is defined as the dis-
tance away from the surface where the inviscid flow properties at
the wall are asymptotically reached. When the flow field is treated
globally, which is typical for predictions based on Navier-Stokes
equations, it is not easy to distinguish between inviscid and vis-
cous portions of a flow field. This is why the distribution of vor-

- 96 -
ticity is used to determine length scales in turbulence models, see
e.g. [108]. In the case of flows with separation the displacement
thickness with its usual definition, resulting from an integration
procedure from the surface to the boundary-layer edge, yields length
scales too large for the use in eddy viscosity turbulence models,
as was shown for two-dimensional flows [112]. These difficulties
are increased for three-dimensional flow-field predictions, since
there only the definition of the displacement thickness can be ex-
tended from two to three dimensions in a straight-forward manner,
see chapter 3 and [ 1] . Another problem is the definition of the
wall-normal distance and the boundary-layer thickness, which are
needed for eddy-viscosity models, in the case of corner flows [115].

In the following another definition for the viscous7layer thickness


is suggested. Consider the surface-tangential velocity component
*1 3
v (x) of a flow past a curved surface, Fig. 8.2 (for the sake of
convenience a two-dimensional flow is treated).

VISCOUS
flow

strongly viscous
flow

Fig. 8.2: Profiles of the tangential velocity component v*1 of a


two-dimensional flow past a curved surface: a) classical
boundary-layer approximation, b) real flow situation
(to be predicted by Navier-Stokes solution).

- 97 -
In first-order boundary-layer theory, see Fig. 8.2, case a) which
is sufficient at high Reynolds numbers in most cases (see chapter 9),
the tangential viscous velocity component v*, (x 3 ) exhibits vanishing
wall-normal derivatives an v*'/(ax 3 )n = an v'/(ax 3 )n (because a" ;
a" (x 3 » at the boundary-layer edge o. The latter usually is de-
fined by the location

(8.34)

Integral thicknesses are defined in terms of local flow variables,


scaled with those of the inviscid-flow solution at the surface, and
integrated from the surface to the edge of the boundary layer. If
the inviscid tangential velocity component v~ is considered, it is
safe to say that, in general, v~ decreases with increasing distance
x 3 from the wall. If inviscid-viscous interaction is taken into
account, the effective thickness ratio of the body will increase

,
due to the displacement effect of the boundary layer, and there-
fore also the ve-component at the displaced surface. This effect is,
except for supercritical transonic flows, usually negligible.

Consider now case b) of Fig. 8.2 where schematically the real velo-
city profile in the neighbourhood of a surface is given. If the ex-
ternal inviscid flow is irrotational, the edge of the boundary layer
can be defined by investigating the magnitude of the tangential co-
ordinate of the vorticity vector (see appendix C, eq. (C.4) with
g-metric) :

(8.35) 0:

w
2
I-
Again only the two-dimensional case is considered.

For high Reynolds numbers the term v~3 in eq. (8.35) - v~3 = 0(')
because of the boundary-layer stretching applied in eq. (8.35) - is
the dominant term. In the frame of boundary-layer theory the boun-
dary-layer edge could be defined therefore as the point where

- 98 -
*1_
v'3 = O. The edge defined by eg. (8.35) lies above this point, be-
cause of the other terms in eg. (8.35). How far the point where
*1
v,) = 0 can be used as approximative criterion for the boundary-
-layer edge in Navier-Stokes solutions cannot yet be decided.

- 99 -
9. On the Influence of Surface Curvature

9.1 Hierarchy of Governing Equations

Whenever boundary-layer flows on curved surfaces are to be predic-


ted, the investigator needs an answer to the question whether the
application of first-order boundary-layer equations is sufficient,
whether second-order theory needs to be considered, or whether a
Navier-Stokes solution, in one form or the other, is required.
Currently, Navier-Stokes solutions are mainly being used to cope
with interaction problems such as separation or near-wake flows,
see also chapter 8.1. Here, with respect to the choice of the
appropriate set of boundary-layer equations surface curvature is
considered to be the dominant factor (see also chapter 7 ). For the
discussion of other higher-order effects, for instance due to vis-
cous displacement, see the textbook by van DYKE [ 6].

While first-order boundary-layer equations are being given in


chapter 2 and second-order boundary-layer equations in chapter 7 ,
the notation "zero-order boundary-layer equations", which appears
now and then in this book, needs some explanation. "Zero-order"
concerns the influence of surface curvature and means that in the
first-order boundary-layer equations surface-curvature terms in the
metric tensor are completely neglected. This represents a good
approximation in the case of only slightly curved surfaces which
exist, e.g., on large portions of a wing. Near the leading edge,
of course, the approximation becomes questionable, see chapter 9.2.
For the prediction of turbulent flows on wings, however, the error
due to the turbulence model and to the assumed location of the
transition laminar-turbulent is certainly much larger. Furthermore,
if leading-edge contamination (for the state of the art see e.g.
POLL [70]) occurs, the ratio of thickness of the turbulent boun-
dary-layer at the leading edge to the smallest radius of surface
curvature there may be so large that second-order theory needs to
be considered. This, however, is meaningless due to the lack of an
adequate turbulence model for that highly three-dimensional and
curved region, so that the prediction is in error anyway. The ad-
vantage of zero-order theory is the simple form of the governing
equations or, at least, of the geometric properties of the sur-

- 100 -
face to be computed at every location (x',x 2 ). There exist quite a
few cases where zero-order theory predicts reasonable results, see
e.g. [42,43,591. For flows past bodies like fuselages, etc. zero-
order theory cannot be applied, of course, since at least the trans-
verse curvature must be considered.

The situation on swept wings is used in the following chapters to


discuss the influence of surface curvature in more detail. The key
parameters of the surface curvature is the magnitude of the curva-
ture itself, and the magnitude of the change of curvature on the
surface. As usual these parameters are compared with the inverse of
the boundary-layer thickness. Among the many possible combinations
of the parameters, the following ones are used for establishshing a
hierarchy of governing equations (hence intermediate classes exist,
and the consideration of terms additional to those given in this
book may improve predictions for certain special cases):

a) Zero-order theory (equations of chapter 2 )

(9.1) 0 (e) «
,
0 ('6) ,
a-metric with Qc (X',X 2 ) - 0

Here e = max (IK,I,IK21) denotes the magnitude of the largest cur-


vature at the point under consideration (see appendix ~, eqs.
A.4' and A.42), 6 is the boundary-layer thickness, and Q c the con-
tour angle (see chapter '0 ).

b) First-order theory (equations of chapter 2 )

(9.2) 0 (e) « 0
a-metric with Qc

c) Second-order theory (equations of chapter 7 , see also [ 31 )

(9.3) 0 (1) < 0 (e) < 0 (~) ,


g-metric with Q c = Q c (x',x 2 )

together with

- '0' -
(9.4) I) o (1) < 0 (D) < 0 (1) ,
IS

and

(9.5)

where D denotes the largest gradient of the surface curvatures along


the surface. For details see chapter 7 and ROBERT [ 3] . ROBERT
does not use the notation "second-order" theory, which, strictly
speaking, should be used for case c I) only.

d) Navier-Stokes equations (equations of chapter 8 )

(9.6) 0 (e) 0 (1 )
IS

0 (D) 0 (1)
IS
g-metric with u Uc (x 1 , x 2 )
c

It is noted that in the above hierarchy of equations it is the re-


presentation of the inviscid (pressure) flow properties which is
successively being improved, and not, except maybe for case d), the
representation of the viscous (and turbulent) flow properties.

- 102 -
9.2 Boundary-Layer Flows at the Leading Edge of
Swept Wings

9.2.1 A Criterion for the Validity of First-Order Theory


for Laminar Incompressible Flows

Consider a swept wing without nacelles and the like mounted on a


fuselage. The regions of highest curvature of the wing are then
found near the leading edge, and, of course, at the root and at the
tip. Here the leading-edge region only is considered, and in the
following a simple approximate criterion is developed which allows
to decide whether first- or second-order boundary-layer theory has
to be employed for the prediction.

In order to arrive at a simple formula the following assumptions


are made:

1. The wing is a cylindrical wing with symmetrical cross section


and at zero angle of attack

2. Surface curvature is negligible in the transformation of the


chordwise derivative (V~'1)

3. Infinite-swept-wing flow

4. Laminar, incompressible flow

In Fig. 9.1 the planform and the cross section of the wing under
consideration are given together with the generally used non-ortho-
gonal (a) and additionally with orthogonal (b) surface-oriented co-
ordinates.The representation of the surface of the wing in the Car-
tesian reference coordinate system in terms of the surface para-
meters is for case a) (all lenghts have been non-dimensionalized
with Lref ) :

- 103 -
l'
X

I
_2_2
X,X

LX,..,,2-L
- X2-5
-

1'
X
Lx' =C05 ~ C
Fig. 9.1 Schematic of a swept wing with non-orthogonal (aI,
and orthogonal (bl surface-oriented coordinates.

- 104 -
x1
(9.7) x1 ' tg "'0 s x2 + c
0
J cos a c (f; 1 ) d f;1

2' 2
x s x
x1
x 3'
cJ
0
sin a
c
( f; 1 ) d f; 1

and for case b)

x- 1
(9.8) x1' tg "'0 s ;C2 + cos 2"'0 cf cos a c (~1) d ~1
o
-1

J
x
x
2'
s x-2 - sin "'0 cos "'0 c cos a c
o

-1

J
x
x3 ' cos "'0 c sin a c (~1) d ~1
o

where the tilde is used here and in the remainder of this chapter
to distinguish between the two sets of coordinates, and not, as
usual, to indicate boundary-layer stretching.

The transformation matrices for case a) and b) follow from the dif-
ferentiation of eqs. (9.7) and (9.8) (see chapter A.4 and also
A.3, appendix A). For case a) one obtains:

(9.9) sin a c (x 1 )
cos a c (x 1 ) c tg "'0 s
N

tg '" sin a (x 1 )
B 0 s o c
N

cos a (x 1 )
sin a (x 1 ) c 0
c
c N

- 105 -
and

(9.10)

tg ~o cos a c (xl) sin a


c
c (N)2 c (N)2 c

tg ~o sin 2ac (xl) tg ~o sin ac (xl) cos ac (xl)


s (N) 2 s (N) 2

cos a
c
N N

where

(9.11) N [1 + t 9 2 ~o Sln
.2a (x 1 )]1/2
c

The corresponding matrices for case b) read:

(9.12 )

cos
2
~o c cos a
c
(~1 ) tg ~o s - cos ~o sin a
c (~') \
c cos a sin a (~, )
- sin c (x, ) s
B ~o cos ~o sin ~o c

c sin ac (~1 ) cos a


- (~, )
cos ~o 0 c

and

- 106 -
(9. '3)

cos ac (~') tg 11>0 cos ac (~') sin ac (~')


c c cos 11>0 c

sin 11>0 cos cos 211>0


B-' S
11>0
S
0

-cos 11>0 sin ac (~') sin 11>0 sin ac (;,) cos ac (;,)

Consider next the relation between the nose radii rN and rN with
respect to the two different surface-oriented coordinate systems,
which requires the relation between two corresponding contour angles
a c and ac • This can only be achieved in an approximate way, the re-
sult, however, turns out be exact for special configurations, in
particular for the leading edge of a wing.

In order to derive the relation, a new Cartesian reference coordi-


nate system (;i') is introduced whose origin coincides with the ori-
ginal (xi') system, but which is rotated by II> with respect to the
original one, see Fig. 9.2. The point P (x" ,~2' = 0) has approxi-
mately the same x'- and ;'- coordinate, if the wing is sufficiently
flat, since the physical coordinates are normalized with the re-
ference lengths Lx' and L;', see Fig. 9.2 (see also chapter '0.'.').
The point p' (x-,' = cos II> x " , x-2' = 0) has the same x-, -coordinate
o ,
as P and, therefore, also approximately the same x -coordinate.
The tangent of the contour angle in P or P' in xl-direction is
given by

(9. '4) tg a c , '


a3 ' / a, ,

and in xl-direction by

(9. '5) tg a c a~'/ a-i' ,


Since x 3' x- 3' and x
-, , cos m x" and since x' = x' in P or P',
"'0
one obtains

- 107 -
x 2'

x
"

Fig. 9.2 Planform of a cylindrical infinite swept wing with two


Cartesian reference coordinate systems (xi') and (xi').

(9.16) tg Qc cos !Po tg Qc

Relation (9.16) is exact at the leading edge and at the trailing


edge of an infinite swept wing, and, obviously, on the apex of a
circular cylinder and on (supersonic) double-wedge profiles. The
derivative of Qc with respect to x 1 can be expressed as follows:

(9.17) Qc'1

Because all derivatives of Q or Qc with respect to or x 3 ,


-3 c
x vanish, the expression (9.17) yields

-108 -
(9.18) «c'l

and one obtains further with the help of the corresponding elements
of the transformation matrices (9.9) and (9.12), after use has been
made of equation (9.16):

cos ljIo
(9.19) «c'l ------~--------~~----~--~~-
(cos «c + cos 2
2~ . 2«~) 3/2 aC'l~
ljIo s~n .c

Taking the limit «c ~ n/2 at the leading edge, and introducing the
physical lengths of the coordinates yields the relation of the cur-
vatures at the leading edge

(9.20) «c'l*

and thus that of the nose radii as well

(9.21) rN

With the geometric properties of the wing given the well-known re-
sult for the thickness of the laminar incompressible boundary-layer
at the attachment line of an infinite swept circular cylinder
(see e.g. ROSENHEAD [10] or SCHLICHTING [11]) is employed for the
derivation of the criterion:

(9.22)

where the gradient of the external (inviscid) velocity component


is always positive. Due to the assumption of an infinite length of
the swept wing, the boundary-layer thickness 6 is constant along the
leading edge. This constant c, is c, ; 3 for 6 0 . 99 and c, ; 4 for
60.9999. If the variables are made dimensionless as follows (the
overbar denotes dimensional variables):

- 109 -
(9.23) ;5

\! \! \!ref

,,*1 -1
x x L-1 L
x ref

rN rN L ref

equation (9.22) becomes

(9.24)

-1
In order to express the velocity gradient ve'l in terms of the gra-
dient v:'l in the non-orthogonal xa-coordinate system, Fig. 9.1 a),
the wing is considered as having no thickness, since the transfor-
mation of the derivative for a real wing is very tedious. With
ac a c - 0 in eqs. (9.9) to (9.13) and using the infinite-swept-
wing assumption a/a x 2 = a/a x
2 = 0 as well as the condition that
the leading edge is a line of symmetry, one obtains:

-1 1
(9.25) v v
e'l e'l

Thus relation (9.24) reads

c, 1
(9.26) <5
IRe
ref Iv!, 1
where the constant c 2 takes care also of the error associated with
the assumption of flatness of the wing.

Now, according to eq. (9.2), first-order boundary-layer theory can


only be used if the boundary-layer thickness is small compared to
the smallest radius of curvature of the surface of the wing (note
that rN represents the maximum radius of curvature at an infinite
-2'
wing with the leading edge parallel to the x -axis) :

- 110 -
(9.27)

Furthermore it is assumed that the following relation holds for


the nose radius of an airfoil [41:

(9.28)
rN ~ (!)2
c c '

where t denotes the maximum thickness of the airfoil. Then the


criterion (9.27) together with eqs. (9.21), (9.26) and (9.28) takes
the following form:

(9.29) 6
- -
rN

Eq. (9.29) contains the usual requirement for fixed geometry, name-
ly that the reference Reynolds number be large, since the boundary-
layer thickness decreases with increasing Reynolds number. Further-
more, the thicker and the larger the wing, the larger the nose ra-
dius and the acceleration at the leading edge, the smaller there-
fore the ratio 6/ r N . The effect of the angle of attack is also in-
corporated in formula (9.29) via /v~'1. The appearance of the sweep
is due to the effective chord c cos ~o for infinite swept wings.
Fig. 9.3 shows, with c 2 cos ~o/c assumed to be unity, that for Rey-
nolds numbers larger than 10 6 first-order boundary-layer theory
for laminar flows is valid for nearly any value of the remaining
parameters.

The criterion (9.29) can only be used as an approximation at general


wings in particular at higher speeds, since it results from rather
crude assumptions. Especially at high angles of attack the boundary
layer may have grown considerably from the attachment line beneath
the leading edge when the highly curved nose· region is reached. In
such cases a more elaborate estimate can be made by following the
procedure discussed in chapter 9.3. A very serious problem, however,

- 111 -
arises if the flow is already turbulent at the leading edge due to
leading-edge contamination [34,70]. No solution for the turbulent
boundary-layer thickness equivalent to that for laminar flows, eq.
(9.22), is known to the authors. Therefore no estimation canbemade
for turbulent flows as to when second-order theory should be con-
sidered, and, more important even, when curvature effects have to
be incorporated into the turbulence model for such three-dimensio-
nal flows.

1
tic
"e~,
,
,,-, D 50 0.05

"""
+ 100 aDS
,~ ~ <> 150 aDs
0.1 t:. 50 0.10
, , 0 100 0.10
"""
.,""'" 150 0.10
~ " '"
I>

"""
I'.. "
50 a15
tom ~"
)(

~ ~ ~~ ~ a 100 0.15
V 150 0.15
~
~
,
'"
~""
, , .....
l"..'
-- r-- ~-

0.001
""" ~"
~ ~ ~~""" ~
..... ..... ,,,...... ~ .......
-~~

"'"'" ~"
~~

"'
"""
.

I I ~~
0.0001
-. ~~. ---t- - ~
- ~
.~

Reref::----..
~

Fig. 9.3: Distribution of the curvature parameter, eq. (9.29), ver-


sus reference Reynolds number with the chordwise velocity
gradient at the leading edge and with the thickness ratio
of the chordwise profile as parameters.

- 112 -
9.2.2 Results for Zero- and First-Order Theory

The definition of zero-order theory was given in chapter 9.1, the


corresponding results are easily obtained from those of first-order
theory by setting the contour angle equal to zero. The surface-me-
tric tensors for first-order theory (a-metrics) of the coordinate
systems, depicted in Fig. 9.1, read for case a)

(9.30) (c) 2 tg 1110 s c cos Q


c
aae
tg 1P0 S C cos Q c

and for case b)

2
(9.31) (cos 1P0 c) o
aae
0

For the flat wing with a c - 0, the metric tensor becomes for case a)

(9.32) (c)2 tg 1110 s c


a'
ae 2
tg 1110 s c (co: 1110
)
It is interesting to note that for the orthogonal coordinate sys-
tem, case b), the surface curvature does not at all affect the sur-
face-metric tensor. Only for the g-metric for x 3 > 0 the contour
angle would appear. Thus zero-order theory has its meaning only
while dealing with non-orthogonal coordinates. Then, however, the
geometrical properties of the wing under consideration simplify
considerably. The number of metric factors k mn , defined in chapter
2., reduces to a certain extent as well.

- 113 -
For the laminar flow near the leading edge of a swept tapered wing
results for zero- and first-order theory are compared with each
other in Fig. 9.4. For a cross section near mid span the location
of separation on the suction side is given for zero-order (a = 0)
1 c
and first-order (a c = a c (x )) theory for several angles of attack
[37]. The experimental results [74] show the existence of a sepa-
ration bubble for angles of attack larger than 5 degrees. The zero-
order-theory predictions locate separation somewhat more downstream
which is in accordance with experience in two-dimensional boundary
layers [29,31]. These results are in remarkable agreement with
those for first-order theory although the flow in the experiment
is very complicated (compressible; flow unstable at xl ~ 0,02;
three-dimensional bubble separation). For a = 15 0 larger separation
zones exist already near the leading edge resulting in a wrong pre-
diction of the inviscid flow field which is documented by the dif-
ference of the predicted and experimental locations of the attach-
ment line in the left portion of Fig. 9.4 [37]. For practical pur-
poses the difference of the predicted separation locations is neg-
ligible. Since bubble transition occurs, the question of bubble
length and initial conditions for the boundary-layer computation
after turbulent re-attachment is much more important [23]. These
initial conditions influence strongly the prediction of the now in
general decelerated turbulent boundary layer and may govern comple-
tely a calculated vortex-sheet separation location near the trai-
ling edge of the wing.

Further away from the leading edge the influence of surface curva-
ture is negligible in laminar-flow calculations [36,37]. A study
of two-dimensional flows by GRUNDMANN and ROBERT [31] indicates
that second-order effects are small even if the stagnation point
on the lower surface of the airfoil is markedly away from the
leading edge.

Thus it may be concluded from investigations of conventional wings


and airfoils (wings, where the region of high curvature is restric-
ted to the very neighbourhood of the leading edge), that zero-order
theory may safely be applied in flow predictions for such wings.
Exceptions are cases with leading-edge contamination, where, pre-
sumably, the problem is the appropriate turbulence model, and stu-

- 114 -
lower- upper side of wing
laminar separation
(sep. bubb/~J

<:> experiment
o
5
o attachment line
experiment

..
0 0
-QO' -0.02 0 0.02 0.0' 0.06
x,

J' 2' 2
X x I X x' = const.

Fig. 9.4: Comparison of predicted and experimental locations of in-


compressible laminar separation (leading-edge separation
bubble) on a swept tapered wing [37].

dies of hydrodynamic stability which become more and more impor-


tant and which require highly accurate predictions of the basic
laminar flow fields.

- 115 -
9.3 Some Remarks on General Cases

Not much can be said regarding flow predictions on general shapes.


The only way to check the validity of first-order theory seems to
be the a-posteriori check: the predicted boundary-layer has to be
compared locally with the smallest radius of curvature of the sur-
face. If the ratio appears critical, only a second-order calcula-
tion can show the real influence of curvature. No a-priori estima-
tes, apart from very crude ones, can be made for corner flows.

Viscous flow in highly curved regions is, in general, no longer


boundary-layer flow. Typical examples are a) flow along sharp con-
cave corners, from which one or several pairs of longitudinal vor-
tices may originate, b) flow at forward- or backward-facing steps,
which induce separation with recirculating flow regions, c) flow
at wing-body junctions, where the three-dimensional separation of
the body boundary layer in front of the leading edge of the root
section of the wing leads to the formation of a vortex (with or
without secondary vortices) with its axis in chordwise direction,
etc .. The reader is refered to the abundant literature on these
and similar topics which, so far, withstand very much any theore-
tical/ numerical treatment. Reasons are the very complicated local
and global interference of the viscous shear layer with the exter-
nal flow, the very small scales of the secondary and tertiary struc-
tures of vortices and vortex sheets, and resulting turbulent pro-
perties. The only means for a successful treatment is seen in the
numerical solution of the Navier-Stokes equations, which is, how-
ever, handicapped by the lack of sufficient computer power and, in
particular, by the lack of understanding of the phenomena in ques-
tion.

- 116 -
10. Samples of Surface-Oriented Coordinate Systems

In this section surface-oriented monoclinic coordinate systems


are considered for both wings and bodies. For the sake of con-
venience the geometries of the configurations considered are
somewhat idealized to remain amenable to an easy analysis. The
concepts shown here are valid also for more complicated aero-
dynamic shapes. In practice, one would like to use a numerical
geometry package in order to determine the traces of the chosen
coordinates on the surface of the wing or body. Such a geometry
package would lend itself to the application of a straight-forward
procedure to determine the metric properties of the surface. Suppose
the mesh on the surface of the wing or body considered be given in
terms of the reference Cartesian coordinates of the nodes formulated
as functions of the chosen Gaussian surface parameters. Then the
metric properties of the surface, required for the integration of
the governing equations, are obtained by simply evaluating the
derivatives of the Cartesian coordinates with respect to the surface
parameters (see appendix A). This approach facilitates the work of
the investigator, once the package works. It will, however, require,
in general, increased computational efforts. The approaches,
discussed in the following chapters, especially the use of contour
angles a (x 1 ,x 2 ) , give more insight into the problems encountered,
c .
and moreover considerably reduce the computational effort needed to
obtain the surface metric properties, since they profit from nearly
analytic relationships. Concerning the application of the surface-
oriented monoclinic coordinate systems [ 1, 3,351, which are
being discussed here, the reader is reminded of the fact that they
are particularly suitable for convex or mildly concave surfaces but
cannot be used for, for example, corners. If such a case is to be
considered, one has to use general coordinates [ 21 and the
governing equations with the metric terms retained in general form
(see also chapter 8).

- 117 -
10.1 Surface-Oriented Wing Coordinate Systems

10.1.1 General Wing Coordinate Systems

Consider a wing with a general planform shape with smooth


leading and trailing edges as shown in Fig. 10.1. If the
leading or the trailing edge exhibits breaks, these have to be
considered explicitly,or they have to be smoothed out for the
purpose of the flow prediction as is often done. In the following
all lengths are made dimensionless by means of a reference length
Lref such as a mean chord length or the like . In all geometrical
3
considerations no boundary-layer stretching is employed for x

The coordinate lines x 2 const. originate in the leading edge and


lie in planes of constant span which are planes parallel to the
1I 3I 1
x x -plane. The coordinate lines x const. originate in the
root section and link stations of constant percent of arc length
in the chordwise direction of the individual cross sections. Note

Leading edge (L. E)


x' =const.
Root

x2 =const.
Trailing edge (T E)
t---------s

x"

Fig. 10.1 General wing planform shape (v 1 , v 2 , ~ and coordinate


1I 2 I
lines are projected into the x , x -plane).

- 118 -
2 2'
that the x -distances are measured along x and not along the
2 1
x -coordinate lines (lines for which x =const.). If the percent
lines are straight lines, x 2 divides equal portions on s and on the
percent lines*).

In the spanwise direction the shape of the wing may be quite


arbitrary, the wing may, in particular, have twist. A typical
cross section of a wing in a plane x 2 ' = constant is depicted in
Fig. 10.2. In each cross section the arc length, measured on the
surface from the leading to the trailing edge is given by

(10.1)

on the upper side, and by

(10.2)

X 3'

l'
XL.E. X 7'

Fig. 10.2 Cross section of the wing, x 2 x 2 ' = constant


(a c is measured in planes x 2 ' constant) •

*) Note that in the frame of this book only surface parameters


are being employed. One could as well choose, for instance,
an angle to define the surface-points.

- 119 -
on the lower side (Fig. 10.2). These lengths are taken to be the
local normalizing lengths for xl (lengths in chordwise direction).
It is, of course, possible to measure these lengths, including the
xl-coordinate, from the attachment line, which divides the flow on
the upper surface from that on the lower surface, and which
therefore naturally separates the corresponding domains of inte-
gration for the flow equations (see discussions in chapter 5).
This, however, would require to compute the metric properties anew
whenever the angle of attack changes.

Let the surface coordinates be such that:

leading edge (LE) : xl


0 :; x 1 :;; 1
(10.3)
trailing edge (TE) : x
1 ~} ,
root 2
(R) : x 2 :;
tip (T) : x
2 ~} 0 :; x 1 ,

with the leading edge given in terms of the Cartesian reference


frame by

2
(10.4) 0, x ) ,

3I 1 2
(10.5) x LE (x oI X )

Using the contour angle

(10.6)

one obtains the relation between the xiI-reference coordinate


system and the xa-surface coordinate system on the upper surface:

J
xl
(10.7) Xul' (x 1 ,x)
2 = x l' (O/X)
LE
2 + Lx1 (x)
2 cosa c (~ 1 ,x 2 )df; 1 I
u 0 U

(10.8)

- 120 -
3' 1 2
(10.9) Xu (x ,x )

Similar expressions follow for the lower surface.

In the case of simple cylindrical or conical wings all cross


sections are identical or similar, and the normalizing length
in chordwise direction can be chosen for simplicity as

l'
(10.10) x TE (x 1 = 1,x 2 ) l'
- x LE (x 1 = O,x 2 )

so that the coordinate xl at the trailing edge will be xl = 1 + £


where £ > 0 represents the difference betweeen the arc length on
the surface and the chord length.

In the following the geometry for only one side of the wing, the
upper side, will be considered. The index u is therefore omitted.

The coordinates eai ' (eq.(A.l0)) of the base vectors -a


a (a= 1,2)
are found by means of equations (10.7) to (10.9):

ax 1 ' 1 2
(10.11 ) 15 11 ' -1- (x 1 ,X 2 ) Lx 1 (X 2 ) cosac(x ,x )
ax

2' ax2' (x 1 ,x 2 )
(10.12) 15 1 - ·0 ,
ax 1

3' 3' 1 2
(10.13) 15 1
ax (x 1 ,x 2 ) Lx 1 (x 2 ) sinac(x ,x )
~

ax 1 ' 1 2 a (x 1 ' (0,x 2 ) ) +


(10.14) 15 12 ' -2- (x ,x ) LE
ax ax 2
x1
+ Lx 1 (x 2 ) J a 1 2 1
- 2 (cosac(E; ,x )) dE; +
ax
0

- 121 -
1 2 1
cosa c (~ ,x ) d~

ax 2 I 1 2
(10.15) - - (x ,x ) -
ax 2

3I 2
(10.16) (x LE (o,x )) +
ax 2

a (sina (t;1,x 2 )) dt;1 +


ax2 c

The coordinates B~' of the third base vector can now be found
following appendix A.3. Then all metric properties and trans-
formation matrices can be calculated as shown in appendix A.

In the following a simple example is given: a conical swept wing


whose planform is sketched in Fig. 10.3. Since all cross sections
1
are similar the angle a c of the contour is a function of x only:
a = a (x 1 ). The normalizing length L 1 becomes with eq. (1 0 . 1 0) :
c c x

(10.17)

cR -

The xiI-coordinates are then related to the surface coordinates


x a (see eqs. (10.7) to (10.9)) as follows:

x1
(10.18 ) X
1I
[c R- (cR-c T ) x 2 ) J
o
2I
(10.19) X

- 122 -
xl
3'
(10.20 ) x [c R-(c R-c T )X21! sl.·na(~l) d~l
<, <,
.
o
The coordinates of the base vectors ~1 and ~2' eqs. (10.11) to
( 10. 16) , become

(10.21)

2'
(10.22) 13 1 0 ,

3' 2
(10.23) 13 1 [c R - (cR-cT)x 1 sina (x 1)
c
1
x
1'
(10.24) 13 2 =tg"'LE s - (cR-c T ) ! cos a (~l)d~l
c
0
2'
(10.25 ) 13 2 = s ,

x':: 0

v'
")(1=1+£
x ::const.
X, =const.
~----------------s----------------~

X l'

Fig. 10.3 Planform shape of a swept conical wing ('" is the


p
local sweep angle of the projection of the
2
x -coordinate into the x
l' , x 2' -plane, *,v 1 , v 2 are

projected into that plane, too)

- 123 -
xl
('0.26) (cR-c T ) J sina c (E; 1) dE; 1
o

and the metric tensor of the surface is

(10.27)

1
x
(10.28) = (s) 2 { 1 + [tgCPLE -
1
---
s
o
J
+ [-
cR-c T
s
Jx
o

(10.29)

In the present example the tangent of the projected local sweep


angle cP (see Fig. 10.3) is
p
1

(10.30) CR:c T J
x

and the tangent of the local elevation angle E of the coordinate


line x' constant

3I 3I
x R -x T
1
(10.31) tg E = tgE (x )
s/coscp
P

Using these two geometrical definitions, and the definition of


Lx' , the components of the metric tensor are rewritten:

(10.32)

- 124 -
s 2
(10.33) a 22 = (COSf/l casE) •
p
The latter expression is the square of the length of the coordi-
nate line x 1 = canst. on the wing surface. This length relates
the x 2-coordinate lines to physical lengths (see chapter 2).
Keeping in mind that

(10.34) ~ ,Ia 22 cos3 ,

where 3 is the local angle between lines of constant x 1 and x 2


on the surface, one gets in this case

(10.35) s
Lx 1 cos lP CaSE cos3 ,
p

and

(10.36) cos3 (x 1 ) COSQ c (x 1 ) sin lP p (x 1 ) CaSE (x 1 ) - sinQc (x 1 ) sinE (x 1 ) •

Because the curvature properties of the surface cannot be dis-


cussed explicitly for this example, this is done for another,
still simpler example, namely a swept cylindrial wing. Here
<Pp = f/l LE and Lx 1 :; c R = c T hold. If the constant chord length
is taken as reference length, Lx1 is equal to one, and the metric
tensor of the surface reads (see also [36]):

(10.37)

The components of the third base vector ~3 resul ts from eqs.. (A. 1 J)
to (A.15):

(10.38)

- 125 -
tg ~LE sin a c (xl)
(10.39)
N

(10.40)
N

with N [ 1 + tg 2 ~LE
.2 a
S1n (x1»)l/2.
c

The curvature tensor of the surface follows from eq. (A.40):

o
(10.41) b
at!
o

From eqs. (A.41 and (A.42) one gets the two principal curvatures
of the surface:

(10.42)
+ tg 2 ~LE da c (xl)
(N)3 dx 1

The directions of the two principle curvatures are with eq. (A.46)

(10.43) dx' o ,

- s

The first represents lines x' = const. that is the x 2-coordinate


lines, and the second lines orthogonal to the x 2-coordinates.
The reader can verify this easily at a flat wing with a c = O.

- '26 -
The contravariant metric tensor (A.48) of the surface of the swept
cylindrical wing becomes:

tg ~LE cos ac (x 1 )
s (N)2

(10.44) aa s

/
tg ~LE cos a c (x 1 )
s (N)2 (s N)2

The components of the mixed variant curvature tensor (A.49) read

2
1 + tg ~LE

(N) 3

tg ~LE (1 + tg 2 ~LE) cos a c (x 1 )


s (N) 3

(10.45)

b1 0
2 '

b 22 0 •

The covariant off-survace metric tensor (g - metric) finally has the


components (see equation (A.57)):

X
3 1 da c (x 1 )
(10.46) 1 - 2 - -~"7"-- +
N dx1

- 127 -
(10.47)

2
(10.48) (s/cos Cj) LE) •

The g-metric thus differs from the a-metric only due to g11.
For x 3 ~ 0, a 11 is recovered.

Note that the kind of construction leading to eqs. (10.7) to


(10.9) can also be used in a similar way to generate a mesh in
two-dimensional or axisymmetric channels with arbitrarily shaped
walls if the spanwise direction is identified with the axial
direction while the chordwise becomes the cross-sectional direction.
The resulting mesh yields lines of constant percent of cross-sectio-
nal height and of constant axial distance. Although this is a sur-
face-oriented mesh, it is no longer a mesh for boundary-layer
computations due to the curved walls which result in a non-ortho-
gonal mesh at the boundaries.

10.1.2 Infinite Swept Wing Coordinate Systems

In these cases, see chapters 5.2.1 and 5.2.2, x 2 is measured along


the leading edge. Then, with Ex1 Ex2 = Eref , particularly

2' 2 1,
(10.49) x cos Cj)LE x , Lx 1 L 2
x

and the examples in [35] and [36] are recovered by setting


s :: COSCj)LE in eqs. (10.37) to (10.48).

The metric tensor for the infinite swept wing surface follows
then from eq. (10.37) with

- 128 -
(10.50)
cosa c (x 1 )

for the non-orthogonal case (chapter 5.2.2.).

For the orthogonal case, chapter 5.2.1, the metric reduces to


that of the unswept wing, ~LE = 0:

(10.51)
( 0 :1
Note that in this case the surface properties no lonqer appear
in the metric tensor. The surface curvature tensor b ae of course
does not become trivial, and the metric mixed variant curvature
tensor and the g-metric are non-trivial as well (see also chapter
9) •

10.2 Surface-Oriented Body Coordinate Systems

Two different monoclinic surface-oriented coordinate systems for


bodies are discussed in the following two sub-chapters. The dif-
1
ference is due to the choice of the longitudinal coordinate x •
One system, the cross-section coordinate system, uses the non-
dimensionalized Cartesian coordinate x 1 ' as x 1 , the other system
employs the longitudinal arc length measured on the surface of the
body as coordinate x 1 While the latter system is advantageous in
the case of flow at an angle of attack of zero degree, both sy-
stems fail near the nose of the body if the angle of attack is
non-zero, and a separate coordinate system has to be introduced
there locally (see e.g. CEBICI et al. [20]), unless an external-
streamline coordinate systems is employed [28,76].

- 129 -
In the third sub-chapter the coordinate system used by BLaTTNER
and ELLIS [15] is briefly sketched which does not need a
separate coordinate system for the nose region, but which needs
much computational work similar to the external streamline
coordinate system. For a body of revolution at an angle of attack
of zero degree this coordinate system coincides with the second
monoclinic system mentioned in chapter 10.2.2. STOCK [82]
employs a coordinate system for the prediction of the flow past
an ellipsoid, which has also the advantage that geometrical and
aerodynamic singularity coincide at the stagnation point. In all
these cases, however, the entire coordinate systems has to be
defined anew for changing angles of attack. ,

10.2.1 Pure Cross-Section Coordinate System

Consider a finite body with a fairly general shape as is shown


in Fig. 10.4. The body is defined in a Cartesian xiI-reference

3'
X

Fig. 10.4 Sketch of a finite body in the Cartesian reference


coordinate system.

- 130 -
frame such that the two poles are linked by the xl '-coordinate
axis. The x 2 -coordinate line on the surface of body, i.e. lines
where xl const., are defined by transverse-frame cuts (cross
sections parallel to the x 2' , x 3' -plane) .

The xl-coordinate lines on the surface of the body intersect with


each of the transverse-frame cuts such that x 2 remains constant
(see Fig. 10.5). The coordinate xl, however, is measured along the
x 1 ' axis.Within each cross section x 2 is measured from the point
3' 2' 3'
on the surface with largest x : (x o ' Xo ), see Fig. 10.5. Note
that the center line of the body need not he straight, neither
has the cross section to be symmetrical, in principle.

The circumferential arc length of each transverse frame Lx2(x1) is


the normalizing length for x 2 , whereas L = L ref of the body is the
normalizing length Lx 1 for x l (Fig. 10.5):

(10.52) x1' L 1
xl Lx 1 = xl
x 2
x
(10.53) x
2' 2' 1
Xo (x ) + Lx2 (x 1 ) J
0 2
cosa c (x 1 ,E;2)dE;2
2

J
x 2
3' x~' (x l ) sina c (x 1 ,E;2)dE;
(10.54) x Lx2 (x 1)
2
0

l' 3'
For a body symmetrical with respect to the x , x -plane, of
course, Xo2' 2 1 ) can be chosen as half of t h e Clrcum-
0, and Lx(X .

ferential arc length (see further below) •

The coordinates 8!' of the base vectors ~a(a 1,2) follow from
eq s . ( 1 0 . 52) to (1 0 . 54) :

(10.55) 8 11 ' L,

2' d 2' 1 d
(10.56) 8 = - 1 (x O (x II + - -
1 dx dx 1

x2
cosa 1 I; 2 ) dE; 2 +L 2 (1)J
(x, x _a1 (COSa (x1,C2»dC2
<, <,
c2 x 0 ax c2

- 131 -
A
X
3'
~=const.
x =
2 II

x'= 1
x'=O

~
L-

® tx 3
'
x ' ::const.
x2;; OJ 1

view A-A

2'
Xo
L 2 (x')=circumferentia[ arc length
x , 2
0c
2
=Q'c (x X )
2
J

Fig. 10.5 Definition of the coordinates on the surface of the


body: a) side view, b) cross section x' = constant.

- 132 -
2
x
(10.57) 3' (L x 2 (x 1 » [ .
s~na 1 E; 2 ) dE; 2 -
(x,
B1 c2

(10.58) 1' 0 ,
B2

2' L 2 (x 1 ) cosa (x 1 ,x 2 )
(10.59) B2 x c2

(10.60) 3' L 2 (x 1 ) sina (x 1 ,x 2 )


B2 x c2

In the same manner as in the case of the wing (chapter 10.1) all
metric properties can now be calculated following appendix A. Note
that, in general, the coordinate system will not be orthogonal.

In the following the metric properties of a prolate ellipsoid are


given as a simple example. Fig. 10.6 shows the sketch of an
ellipsoid. The x 1-coordinate is normalized with Lx1 = L = Lref •
The normalizing length for x 2 'in each transverse frame is
Lx 2 = Lx 2(x 1 ) = wr(x 1 ), where w instead of 2w appears, because
due to symmetry only one half of the body has to be considered.

The local radius r is (Fig. 10.5) with D

(10.61)

and the contour angle

(10.62) wx 2

The relation between the surface-oriented coordinate system and


the Cartesian reference system equations (10.52) to (10.54) then

- 133 -
X3' x T =const
X2 = "

xT=O x T= 1
~f------- L -------~

X 3' 2
X=O

view A-A

/ xo2 =O
2' xJ'(x T) =r(x~
X o
L 2 (XT) =Trr(x~
)(

Fig. 10.6 Definition of the surface-oriented coordinate system


for a prolate ellipsoid:
a) side view, b) cross section x 1 = constant.

- 134 -
reads

1
(10.63) x 1 ' = Lx 1 x1 = Lx

(10.64) x 2' r (x 1) sin (lTX 2 ) ,

(10.65) x3 ' r (x 1 ) cos (lTX 2 ) .


The components of the base vectors ~a are as follows:

(10.66) 1' L
111

1 2
(10.67) 2'
81
(0)
"2 r (x 1)
(1 - 2x 1 ) sin (lTX 2 )

3' 1 2
( 1 _ 2x 1 ) cos (lTX 2 )
(0)
(10.68) 81 ,
"2 r (x 1 )

(10.69 ) 1'
82 0 ,

2' lTr(x 1 ) cos (lTX 2 ) ,


(10.70) 82

(10.71) 3'
82 _ lTr (x 1) sin (lTX 2 ) .

The metric tensor then reads (underlined result for general axi-
symmetric body):
(L)2 + lQl2 (1_2x1)2
4 x 1 (1-x 1 ) o
= (L)2 + (r(x 1 ) ,1)2
(10.72)
( IT 0) 2x 1 ( 1-x 1 )
o = (J~x2 (x 1) ) 2

The off-diagonal coordinates a 1 2 = a21 are zero which proves that


this surface-oriented coordinate system is orthogonal as was ex-
pected. The singular behaviour of aall due to this choice of coor-
dinate system will be discussed later.

- 135 -
The components of the third base vector ~3' which is normal to
the surface, are determined with eqs. (A.13) to (A.1S):

(10.73) a31 ' - (1 - 2x 1 )


M

2L/X 1 (l-x 1 ) sin(1Tx 2 )


(10.74 ) a32' M 0

2L/X 1 (1-x 1 ) cos (1TX 2 )


(10.7S) a3'
3 M 0

with

The coordinates of the curvature tensor of the surface follow


from eq. (A.40):

o
(10.76 )
- 2 !1T) 2 Lx 1 ( 1--x 1 )
M

The principal curvatures of the surface, given by eqs. (A.41)


and (A.42), read:

(10.77) -2L

and

(10.78)
-2L
M(D)2

At x = 0,5 the principal curvatures become:

20 2
and
- (L)2 o

- 136 -
and for xl o and xl 1:

2L 2L
and
- (0) 2 - (0)2

The reader may verify this by himself. It should, however, be


mentioned that the principal curvatures are measured in planes
perpendicular to the surface at the pOint considered. The direc-
tions of the principal curvatures on the entire body are
xl = const. and x 2 = const., which is obvious. The coordinates
of the contravariant metric tensor, eg. (A.48), become

(10.79) aafl
o

and these cf the mixed variant curvature tensor, eg. (A.49), read

o
(10.80)
-2L
o (0)2 M

As mentioned in appendix A.l1 these components are egual to Kl and


K2 for orthogonal coordinate systems (see egs.(10.77) and (10.78)).
Finally the g-metric tensor, eg. (A. 57), reads:

31 2
(l-x b 1 ) all
(10.81)
o

The present contravariant formulatioll exhibits two singularities


due to the choice of the pure cross-sectional coordinate system.
In order to see this consider the metric tensor, eg. (10.72) at
the locations xl = 0 and xl = 1, i.e. at the poles of the prolate
ellipsoid. There the coordinate all would diverge leading to

- 137 -
v1 +0, while the coordinate a 22 remains finite, but with
a 22 + 0 inferringv 2 + Note, however, that these singularities
00.

are present also in the classical physical formulations (WANG [86] ,


CEBECI [17]) if one sticks to the coordinate system discussed here.
The principal curvatures, eqs. (10.77) and (10.78), are, of course,
regular.

Due to the discussed singularities difficulties arise in the


solution of the boundary-layer equations in the neighbourhood of
the nose of the ellipsoid. Two cases have to be distinguished:

1. In the case of the flow with an angle of attack of zero degree


the stagnation point coincides with the nose of the ellipsoid.
Then the stagnation point solution can only be computed if either
the coordinate system from chapter 4 (see especially chapter 4.3)
is used, or if another coordinate system is, at least locally,
introduced. This can be, for example, a coordinate system based
on inviscid streamlines. Once initial data are created in the
neighbourhood of the stagnation point at the nose, the remaining
attached flow field can be predicted using the presented coordinate
system.

2. In the case of the flow at an angle of attack the stagnation


point has moved away from the nose to x 1 > 0, and no difficulties
s
arise in creating the stagnation point solution there. However,
1 1
at least in the domain 0 < x < Xs a region with initially unknown
location and extent exists where the flow is directed in the
negative x 1 -direction. Therefore, similar to the case of the
flow around the leading edge of a wing at an angle of attack,
separate computational procedures have to be introduced for the
nose regions and the remainder of the body. Again, an inviscid-
streamline-oriented coordinate system could be used in the vicin-
ity of the nose. However, depending on the angle of attack the
inviscid streamlines are different for each flow case, and there-
fore the metric of the body would have to be determined for each
flow case anew although only for a small domain of the body sur-
face. (This is not necessary for the coordinate system studied
recently by CEBECI et al. [19,20]. These authors employ a non-
singular coordinate transformation for the nose region, further-
more they use an integration scheme which follows the local

- 138 -
streamlines in surfaces parallel to the surface of the considered
body ("characteristic box scheme")). Then an interpolation scheme
has, in general, to be employed at the common boundary of the two
different computational domains in order to provide initial data
for the prediction of the flow on the remainder of the body. Only
in the case where the stagnation point moves to large xl, it
might be useful to employ flow dependent inviscid-streamline
coordinates for the entire body, as has been done by e.g.
GEISSLER [28) , SCH(5NAUER et al. [77) and SCHNEIDER [76).

In the case of a flow at small angles of attack, or for slender


bodies at even large angles of attack the stagnation point is
located close to the nose of the body. If the main interest is
then in the flow further away from the nose, one is able to em-
ploy approximations for the initial conditions, since, due to the
parabolic character of the boundary-layer solution the error in-
troduced decays rapidly downstream. HIRSH and CEBECI [39) used
an azimuthal attachment-line solution at xl x! to initiate the
integration of the boundary-layer equations. In some unpublished
work the present second author used values at xl x! + fixl as
initial data, which were obtained by interpolating profiles predic-
ted in the wind- and leeward planes of symmetry.

10.2.2 Pure Surface-Oriented Coordinate System

A pure surface-oriented coordinate system is depicted in Fig. 10.7.


In this system the coordinate a 11 of the metric tensor of the
surface is always regular, but a 22 ~ 0 for xl ~ 0 and xl ~ 1,
2
leading to a singular behaviour of v . Therefore the problems
described in chapter 10.2.1 still exist. In order to avoid
warped cross sections xl = const. this system should be used only
on bodies with a high degree of symmetry, e.g. bodies of revolu-
tion.

The xl-coordinate follows the surface of the body and is norma-


lized with its total arc-lenghth measured from the windward pole

- 139 -
®
r
X3' A

l'
X

L
A

... 2'
X

(Xc (x 2) = 7Tx 2
2
view A-A
Fig. 10.7 Definition of pure surface-oriented coordinates on a
body of revolution:
a) side view, b) cross section x' = constant.

- 140 -
(x' = 0) to the leeward pole (x' ,). The local radius r(x')
is given by

('0.82) r(x' ) sina


c, (F; , ) dF; ,
where a
c, is the contour angle with respect to the xl-coordinate.

In general, similarly to the computation of Lx1 , the angle ~C1 has


to be determined numerically. Note that in the case of body ~om­
binations, e.g. sphere-cylinder or ogive- cylinder, the surface
at the junctions has to be smoothed in order to avoid a jump in
curvature. To date almost nothing is known about the behaviour
of the boundary layer at such curvature jumps. If the curvature
jump induces a sharp pressure drop and consequent increase,
separation may locally occur.

The relations between the coordinate system x j and the reference


system x i' are (Fig. 10.7)

x'
(, 0.83) x1 ' Lx' J cosa
c,
' ,
IF; )dF;
0
x2
('0.84) x2 ' Lx 2 (x') J
cosa c (F; 2 )dF; 2
2
0
x' x2
1'0.85) x 3' Lx' J sina
c, "
(F; )dF; - L 21x )
x ,J sina
c2
1F;2)dF;2
0 0

The normalizing length Lx2 is half the circumferential length:

('0.86)
,
trr(x ) .

The contour angle a c with respect to the coordinate x 2 is


2

('0.87) trX 2

- 141 -
2
so that the integrations with respect to x can be easily per-
formed in eqs. (10.84) and (10.85).

The components of the base vectors ~a read:

1I 1
(10.88) 13 1 = Lx 1 cos a (x),
c1

2I 1 2
(10.89) 13 1 = Lx1 sina c (x) sina c (x)
1 2

3I 1 2
(10.90) 13 = L sina (x) cos a (x)
1 x1 c1 c2

1I
(10.91) 13 2 =0 ,

2I 1 2
(10.92) 13 = L 2 (x ) cosa (x) ,
2 x c2

3I 1 2
(10.93) 13 = -L 2(x ) sina (x)
2 x c2

The metric tensor is as follows:

(10.94)

which shows that a 11 is a constant for this kind of coordinates,


and a 22 is zero at both the nose and base poles. In the case of
an angle of attack of zero degree the stagnation point coincides
with the nose point and the flow is axially symmetric. Therefore
with a 11s = a 22s =(Lx 1) the stagnation point solution (chapter 4)
can be obtained without difficulty. However, if the angle of
attack is non-zero, the same difficulty arises as was discussed
in chapter 10.2.1.

The components of the third base vector ~3 are as follows:

(10.95)

- 142 -
2' (x 1) sina (x 2 )
(10.96) 63 cosa
c1 c2

3'
(10.97 ) 63 cosa (x 1) cosa (x 2 )
c1 c2

The curvature tensor of the surface is:

1
da (x)
c1
L 1
x o
(10.98) baB

o cos

and the mixed variant curvature tensor:

(10.99)
o

Since the coordinate system is orthogonal the principal curva-


1 2
tures are K1 = 6 1 , and K2 = B2 , which the reader may easily
verify.

10.3 Coordinate System of BLOTTNER and ELLIS

In this paragraph the coordinate system used by BLOTTNER and


ELLIS is sketched only briefly for the case of non-zero angle
of attack. For further details the reader is referred to the
original paper [15). Fig. 10.8 shows the side-view of the traces
of the coordinate system on an sphere at an angle of attack. The
disadvantage common to both systems discussed in the preceding

- 143 -
paragraph is the singularity with respect to v 2 and x 2 at the
windward pole of the body. This can be removed for flow at an
angle of attack if the origin of the coordinate system is moved
with the stagnation pOint, see Fig. 10.8.

X'=o

s X "

Fig. 10.8 Sketch of the coordinate system of BLOTTNER and ELLIS


for a sphere (coordinates not normalized).

The coordinate system shown in Fig. 10.8 is an orthogonal curvi-


linear one. One family of surface coordinates (x 2 = const.) is
generated by the intersection of a plane, turned around the
x 1 '-axis, with the body. The other family of coordinates
(x' = const.)
is obtained numerically by iteration based on the
condition that the lines of coordinate xl = const. and x 2 = con st.
are normal to each other. The procedure thus starts from lines
x 2 = const. and finds for prescribed ~xl the lines xl = const.
Although it may be too costly to define such a coordinate system
for the entire body, the system might be quite useful for the

- '44 -
nose region of a body since the lines x 2 = const. resemble
streamlines close to the stagnation points. Since the origin of
the coordinate system moves with the stagnation point it has the
same disadvantage as the streamline-coordinate system: for each
angle of attack the coordinates and metrics have to be computed
anew. Note that since x 2 , which is an angle, is not defined at
the stagnation point, BLOTTNER and ELLIS assume that the surface
is locally planar there, in order to obtain the stagnation point
flow.

- 145 -
11. Conditions of Compatibility at the Body Surface

The conditions of compatibility at the impermeable body surface,


discussed in this section, are an extension of those for two-
dimensional incompressible boundary layers (SCHLICHTING [11]).
These conditions relate flow variables and their derivatives at
the body surface with each other. Such relationships are important
not only for the understanding of the flow in the immediate neigh-
bourhood of walls but also for series expansions of flow variables
about a point on a surface. Here the continuity, Navier-Stokes
and energy equations, as given in chapter 8.1 for locally mono-
clinic coordinates are considered for compressible flows. Consist-
ent with the remainder of the book the compatibility conditions
will be given for steady flows in terms of the surface metric
tensor aaa for locally monoclinic coordinates.

The no-slip conditions at the wall, usually adopted for viscous


flows, read

i
(11.1) 0: v o i = 1,2,3.

Hence the following relations result at x 3 0:

i i i
(11.2) v 't 0, v 'a 0, v , aa 0, etc.

i = 1,2,3 a,B 1,2.

These conditions reduce considerably the governing partial differ-


ential equations. The continuity equation (8.9) reduces at the
wall to:

3
( 11 .3) P't + P v , 3 o.

Hence, for steady flows one obtains at x 3 0:

3
(11.4) v ,3 o.

- 146 -
Because it is needed later the continuity equation (8.9) is used
to derive a condition of compatibility for the second derivative
v 3 '33 by differentiating it with respect to x 3

(11 .5)

a=1,2.

At the wall this equation yields the condition of compatibility


of v 3 '33 with the remaining flow variables and their derivatives:

3 1 2 1 1 1
(11. 6) v '33 + v '13 + v '23 + v '3 [p P'l + 2g .g'l] +

At the wall the convective terms of the Navier-Stokes equations


(8.10) and (8.11) vanish, and the covariant derivative, eqs.
(8.15) and (8.16), of the stress tensor Tij, eqs. (8.13) and
(8.14), vanishes therefore as well:

(11. 7) Tajlj = 0, o a = 1,2


j 1,2,3.

At the surface of the body one finally obtains for TQjlj

(11.8)

a, fl,y 1,2.

- 147 -
'j
And for T3 J j one obtains

(11. 9)

II 3
+ ~'II v '3 + ~ v '33 = 0 ,

where a,II,~ = 1,2 and j = 1,2,3. Equations (11.8) and (11.9)


represent the compatibility conditions derived from the Navier-
Stokes equations in terms of the g-metric. At the surface,
x 3 = 0, the three relations (11.8) and (11.9) yield the following
expressions, now in terms of the surface metric aaa:

xl-momentum equation

(11.10) O=k 16 P'1 + k17 P'2 + ~'3 v1, 3 + ~ v 1 '33 +

+ .!! v 1 , 3 (- 3 a 22 b 11 + 4 a 12 b 12 - all b 22 ) +
a
2
+ 2 .!!
a v ' 3 (a 12 b 22 - a 22 b 12 )·

x 2 -momentum equation

(11.11)

x 3 -momentum equation

(11.12) o

1 1 2 1 1
+ v '3 kOl k 01 '1 + v '3 kOl k 01 '2 f -

- 148 -
where the second viscosity coefficient A has been replaced by the
bulk viscosity ~' = A + i
~ which, in general, is set to zero.
Equation (11.6) can be used to eliminate v 3 '33 in equation (11.12):

1 2
(11.13) o ~' 1
p (P'1 v '3 + P'2 v '3)

1 1 1 21 1
IL [v , 13 + v , 3 ( ko 1 kO 1 ' 1 + 3" p p, 1 + it ~, 1 ) +

2 2 1 2 1 + 1
+ v '23 + v '3 k"- kO 1 ' 2 + 3" p p, 2 ~ ~, 2) ]
01

This equation relates the pressure gradient normal to the surface


P'3 to quantities which - except for the bulk viscosity ~' - can
be determined experimentally, and thus may serve to estimate the
magnitude of P'3'

The energy equation in terms of the temperature T for thermally


ideal gases eq. (8.12) reduces for steady flows to

(11.14) o

This compatibility condition is difficult to evaluate since


gradients of the temperature in the directions tangential to the
surface are involved. Note, by the way, that derivatives of the
transport properties ~ and k with respect to x 3 contain, in
general, variations with temperature and with turbulence properties
(since in turbulent flows IL and k are taken as effective transport
properties) •

At the surface, x 3 = 0, the compatibility condition derived from


the energy equation reads:

(11.15) 1 2 1 2 2 2
~ [k 41 (v '3) + k42 v '3 v '3 + k43 (v '3) ] +

- 149 -
- .la k {T ' 1

Equations (11.6), (11. 10) to (11. 13) and (11.15) represent the
compatibility conditions at the wall as derived from the governing
partial differential equations. They are given in terms of "primi-
i
tive variables" p,p,T,v . Often it is useful to replace the gradi-
ents of velocity components by terms of wall shear stresses since
these can be measured directly while it is more difficult to mea-
sure velocity profiles near walls.

The complete stress tensor T ij for laminar flows eqs. (8.13) and
(8.14) reduces at the wall to the following expression:

1
(11.16) k16 P k17 P IJ. v '3

2
k26 P k27 P IJ. v ,3

1 2
IJ. v ,3 IJ. v ,3 - P

Note that the wall shear stress tensor is symmetric as it should


be, because k26 = k 17 . The wall shear stresses T are therefore
given by the following formulae where the conventionally used
minus sign has been introduced

3 a3 3a a
(11.17) x 0: T T - IJ. v '3

These relations then serve to replace the derivatives of the


velocity components in the above conditions of compatibility at
the surface:

(11.18) a a3
v '3 T

- 150 -
and therefore

(11.19) a a3 _.1. a3
v , 313 ~ (T '13 ~ ~'13 T )

a,13 1 ,2.

Using equations (11.18) and (11.19) together with the boundary-


layer non-dimensionalizing and stretching procedures (2.6a),
(2.7), (2.9) to (2.111 (see also appendix D) the conditions of
compatibility (11.6), (11.10) to (11.13) and (11.15) yield the
followig relations at the surface x 3 =0 (all stretched quanti-
ties are characterized by tildes, the boundary-layer stretching is
not applied to the components baS of the.curvature tensor):

Equation of continuity (11.6)

~3 ~~ ~13 ~23
(11.20 ) I.1V'33- T '1- T '2

_ ~13 (.1. 1 k
p P'l + kOl 01'1

x 1 -momen t urn equatlon


. (11.10)

(11.21)
~ ~13 1 ~~
0 k16 P'l + k17 P'2 T + ~ v '33 +
1.1 ~'3

-1/2 [~13
+ Re ref (3 a 22 b 11 - 4 a 12 b 12 + all b 22 ) +
a

~23
+ 2 T (a 22 b 12 - a 12 b 22 )]

x 2 -momentum equation (11.11)

(11.22)

- 151 -
x 3 -momentum equation (11.12)

(11.23) o P, '"3 + Re -1 j,l' ) ['"T 1 3 , 1 + ",23


{ (-31 + ~ , 2 +
ref T

",13 "'23 4 ,"'3 "''''}


+ ~ (j,l'1 T + j,l'2 T ) - (3 j,l + j,l ) v '33 •

Modified x 3-momentum equation (11.13)

{"'13 ",23
(11.24) 0 P'3'" - Re -1
ref T , 1 + T ,2 +

4 j,l'
+ "'13 [_1_
p P , 1 (3 + ~) j,l j,l, II +
T +
kOl kO l' 1

+
"'23
T
1
[r k 01 '2 + -p P , 2 (i3 + ~')
j,l j,l j,l, 2l }
01

Energy equation (11.15)

(11.25) o

+ a pr -1
ref
{ -1
Re ref [a 22 (k T'l)'l + all (k T'2)'2 -

- 152 -
The form with stretching of the conditions of compatibility at
the surface clearly exhibits the relative orders of magnitude of
the different terms for viscous shear flows. The compatibility
conditions can, certainly, be formulated in different ways but
the given formulations seem to be most convenient for further
investigations (see, for example, the next section).

In order to give an example for the compatibility conditions, the


flow about a circular cylinder is considered. The axis of the
cylinder, which is sketched in Figure 11.1, coincides with the
x 2 '-axis of the Cartesian reference coordinate system.The radius
of the cylinder is rOo

3'
X X 3'

X l'
Fig. 11.1 Definition of coordinates and of
the geometry of a cylinder •

The components of the base vectors on the surface are (see appendix
A.3) :
1 2' 3' 1
(11.26) 13 11 ' rO cos x 13 1 0 , 13 1 - rO sin x

2' 3'
13 21 ' 0 13 2 1 , 13 2 0

1' sin x 1 2' , 3' 1


13 3 13 3 0 13 3 cos x

- 153 -
The metric tensor of the surface a reads (see appendix A.G)
as

(11.27)

and the curvature tensor baS (see appendix A.7):

(11.28 )

:1
The metric factors k .. reduce to the following non-zero terms:
1)

(11.29 ) - 1

Thus one obtains the following relations for the compatibility


conditions (11.20) to (11.25):

",23 ",13 )
(11.30) o ",3 "''''
~v'33-T
",13
'1 - T '2 - T (1p P'1 - ~ ~'1

(11.31) 0 (r O) -2 P'1 - ~'3


'" "'13 +
T ~
1 "''''
v '33 -
~

3 ",13
Re -1/2
ref T
rO

",23 2 -1 ",23
(11.32) 0 - p, 2 - ~
~'3 '" T + ~ v "'''' -
'33
Re ref
rO
T

- 154 -
o 1 !!:. ') ["'13 ",23
(11.33) P'3'" ref {(-3 + Il
+ Re -1 1: '1 + 1: '2

",23 "'13 ",23 4 "'3 "''''}


+ Il, 2 1: ) 1 + Il (Il, 1 '[ + Il, 2 '[ ) (1
I
Il + Il ) v '33 '

-1
p, '"3 - Re ref {"'13
(11.34 ) 0 "'23
'[ ,1 + 1: '2 +

+
"'13
1: [ (i +
Il')
- Il, 11 +
3 Il P P'1 Il

"'23 [ (i !!:.') 1
+ 1:
3
+
Il P P'2 - i1 Il, 21 }

(11.35) 0

+ Re -1/2
ref

By converting the contravariant quantities into physical quantities


the compatibility conditions derived from the governing equations
usually given in literature are obtained. By taking the limit of
Re ref ~ ~ the conditions of compatibility valid in the frame of
boundary-layer theory are found.

- 155 -
12. Local Topology of Three-Dimensional Separation and
Attachment Lines

12.1 General Remarks

In contrast to the situation in two-dimensional flow the separation


of three-dimensional boundary-layers is characterized by the fact
that, in general, the wall shear stress does not vanish along the
separation line, except in singular points. Usually a distinctive
pattern of limiting wall streamlines, i.e. skin friction lines,
can be realized near separation lines, namely that the skin fric-
tion lines converge towards the separation line from both sides.
Here the separation line itself is considered to be a skin fric-
tion line, which, however, cannot be proved in the frame of local
topology. It is noted that the convergence of the skin friction
lines near the separation line may be a necessary but not suffi-
cient condition for the occurence of separation [102], see for
example the attached flow near the leeward symmetry line of a blunt
body at small angle of attack. The separation line in three-dimen-
sional flow is the location along which the boundary-layer de-
taches from the wall, forming a dividing surface which separates
the converging boundary-layers on each side of the separation line.

In early investigations (see, for example, MASKELL [ 98], OSWA-


TITSCH [100], LIGHTHILL [ 57], EICHELBRENNER [ 25]) lines of three-
dimensional separation are considered as singular skin friction
lines in the sense that they separate, for instance on a body at
angle of attack, the flow coming from the nose of the body from
that from the rear of the body. More recent work shows that, at
least on a body at angle of attack, the separation lines, which are
feeding the pair of lees ide vortices, are not separating the flow
from the front of the body from that from the rear (see e.g. WANG
[103], HAN and PATEL [ 91], MEIER and KREPLIN 99] and KREPLIN et
al. [ 51). Thus the lines of three-dimensional separation are not
singular skin-friction lines in the above-mentioned classical sense.
Note that PEAKE and TOBAK [102) give a profound review on past and
more recent developments concerning the concepts of three-dimen-
sional separating flows.

- 156 -
The converse of separation lines are attachment lines, which occur,
for example, near or at the leading edge of wings, or in the sym-
metry planes of bodies, and which originate from the stagnation
point, a nodal point of attachment in general. It is typical for
such attachment lines that the density of both, external inviscid-
-flow streamlines and limiting wall streamlines, is extremely high,
and that the flows exhibit a strong divergence. In contrast to such
simultaneous behaviour the external inviscid flow at separation
lines does not usually reflect immediately the presence of three-
-dimensional separation. An important special case occurs if the
flow is symmetrical with respect to the attachment line. Then the
attachment-line boundary-layer is a quasi-two-dimensional one (see
chapter 5.1). In general, however, the leading-edge attachment line
on wings is slightly curved, and therefore the inviscid and the vis-
cous attachment lines differ only slightly as well. Note that cur-
vature means here and in the following the local curvature within
the plane tangential to the body at the point under consideration.

In the following local properties of separation and attachment


lines are discussed. The basic assumption is the regularity of the
wall shear stress vector in the region where separation or attach-
ment occurs. The results contribute to an understanding of the re-
sults of both theoretical and experimental studies. No attempt is
being made to tackle the problems associated with the global topo-
logy of such flows. Here the reader is refered to the literature,
e.g. LEGENDRE [ 94, 95, 96, 97], BROWN and STEWARTSON [90], KRAE-
MER [ 93], PEAKE [101], CEBECI et al. [ 20], and the literature
which has already been mentioned. It is noted that the Agardograph
by PEAKE and TOBAK is the most recent review paper. The following
discussion follows closely that of HIRSCHEL and KORDULLA [ 92]
which has been extended to attachment line flows by SCHWAMBORN [ 78].

- 157 -
12.2 Local Topology of Separation Lines [ 92]

In Fig. 12.1 part of the surface of a body with the surface-orien-


ted non-orthogonal ~i coordinate system is sketched together with
a separation line as it can be observed by surface flow visuali-
sation techniques.

3'
X

S1=const., £3= 0
2'
X

X l'
Fig. 12.1 Schematic of coordinates used for the study of separa-
tion lines (~i: surface-oriented non-orthogonal; xi:
surface-oriented locally orthogonal; xi: local Carte-
sian coordinates) •

For the investigation a Taylor series expansion of the velocity


coordinates is used with P (~1, ~2, ~3 = 0) as the local origin.
o
Po is located on the separation l~ne, and a new orthogonal coordi-
nate system on the surface, the x1
-system, is introduced locally
such that its origin coincides with Po and such that the x 1 -axis
is normal and the x 2 -axis tangential to the separation line at Po.
The x 3 -axis coincides with the rectilinear, surface-normal ~3-axis
of the original surface-oriented coordinate system. The Taylor

- 158 -
series expansion can be performed using the surface-compatibility
conditions of chapter 11 and the transformation and series expan-
sion rules provided in the appendix. Because of the large effort
involved if the curvilinear, locally orthogonal xi-system is used,
a Cartesian xi-system is introduced (see Fig. 12.1), which has the
same properties as defined above. Its xa-plane, however, is the
plane which is tangential to the surface at Po' Therefore the vali-
dity of the results of the investigation is restricted to the im-
mediate neighbourhood of the separation line. It is then consis-
tent to take into consideration only the linear terms of the series
expansion. Hence the series expansion for the velocity reads (note
that because of the Cartesian coordinates the velocity and wall
shear stress coordinates have physically correct dimensions) :

(12.1) (x i ) = - {1- [, a3 x 3 +
J.1

a3 J.1, 1
+ ,a3) x 1 x 3 +
h'1 J.1

a3 J.1'2 ,a3) 2 3
+ (, , 2 x x
J.1

1 13
+
23
"2 +
p, 1
(- -)
J.1'1 , 13 +
2J.1 £t'1 p J.1

+
P'2
(-
P
- J.1'2
-)
J.1
,23] (x3) 2 + ••• }p
i 1 ,2, 3
0
a 1 ,2 •
The wall shear stress coordinates and the derivatives of p, p and
J.1 are obtained by transforming the values measured or computed in
the original surface-oriented ~i-coordinates. Included in eqs.
(12.1) and (12.2) are variations due to compressibility or to a
given wall-temperature distribution. Because of the appearance of
J.1'3 in eq. (12.1) knowledge of turbulence properties is required
for turbulent flows.

- 159 -
Eqs. (12.1) and (12.2) were used first by OSWATITSCH [100) in order
to study three-dimensional separation at singular points on the
surface, where the total wall shear stress vanishes. For two-dimen-
sional separating flows the separating streamline was shown to
leave the surface at a finite angle A [100]:

13
3 1:'1
(12.3) tan A
P'1

In three-dimensional flows a limiting streamline leaves the sur-


face only at the singular points mentioned [100), but not along the
separation line, in general. This statement is proved in the follo-
wing, where for the sake of convenience laminar incompressible
flow is assumed. For the discussion the direction of the separating
limiting wall streamline is approximated by that of the separation
line itself. Then, according to Fig. 12.2, the angle A2 , under which
a streamline is assumed to leave the surface, is

(12.4) tan A2

streamline assumed
to leave the surface

separation line

Fig. 12.2 Sketch of streamline assumed to leave the surface.

- 160 -
From egs. (12.1) and (12.2) with x 1 o the following relation is
obtained:

1
(12.5) tan >'2 - 2" 23 + 23 2 1 x3
T T'2 x - 2" P'2 P
o

With tan >'2 ~ x 3 /x 2 for x 3 ~ 0 and x 2 ~ 0 eg. (12.5) can be re-


arranged to yield:

13 T23
+ T'1 + 2 iT
(12.6) tan >'2

No meaningful result is obtained for >'2 unless T23 = 0, which Ip


means that - because of T13 = 0 by definition of the 800rdinate
system - Po is a singular point. This follows from the definition
of the skin-friction line, and the separation line is considered
to be such a line, see Fig. 12.2:

23
T
(12.7) tan 1/I w
""""i3
T

Hence it is concluded that the separation line cannot leave the


surface except at singular points. It is noted that OSWATITSCH's
result, eg. (12.3), is found for x 2 # 0 from eg. (12.6), if the
flow is two-dimensional and in x 2-direction (T 13 = 0) and if T23 0
occurs at Po'

Consider next the separating stream surface whose intersection


with the surface of the body is the separation line, see Fig. 12.3.
It is possible to estimate the angle >'1 of the separating surface
with respect to the x 1-direction at a point P of the separation
. 0 13
line. At P the separating surface is defined by T = 0 and for
3 0 1
x # 0 by v = 0, which - again for laminar incompressible flow -
i
leads to the following expression if the expansion of v 1 (x) in
eg. (12.1) is employed:

- 161 -
Fig. 12.3 Sketch of a separation line and the separating stream
surface emanating thereof.

(12.8)
2 " 131
tan A = - - -
I
1 P'l Po

This result is, in particular, valid for an infinite swept cylinder


where the x 2 -axis coincides with a generator of the surface of the
cylinder (see also following investigations).

The following considerations are confined completely to the limi-


ting surface flow. Therefore only the tangential wall shear stress
components, a3 are of interest. These are obtained by differen-
tiating eq. (12.1) with respect to the wall-normal direction x 3 :

(12.9) , a3 (x 1 ,x 2 ,x 3 - 0) =

a3 a3 Il, 1
+ - ,a3) x 1 +
T
IT'l Il

a3 1l'2 ,a3) 2
+ ("2 Il x + •.• lp a = 1,2
o

- 162 -
In Fig. 12.4 both, the pattern of the skin-friction lines as ob-
served in experiment, and the distribution of the skin friction in
the neighbourhood of the separation line are given. In part a) the
region to the left of the separation line is called forward flow
region, that to the right of it backward flow region.

@ 1-r:I-minimum line

ge

Fig. 12.4 Schematic of a) separation and 1.1- minimum line, and


of b) distributions of the wall shear stress with re-
spect to the x 1 -direction (ge denotes the inviscid ex-
ternal streamline).

Such designations apply locally only because the flow on both sides
may originate from the forward stagnation point. Note that the di-
rection of the external inviscid streamline (ge) may differ con-
siderably from that of the separation line. In part b) typical dis-
tributions of the wall shear stress near a separation line are

- 163 -
sketched. It is seen that the magnitude of the total wall shear
stress 1,1 exhibits a - relative - minimum which is observed in ex-
periments, see e.g. 51], and often interpreted as location of se-
paration. In order to prove that the distribution of 1,1 has an ex-
tremum not at, but near a separation line, consider the derivative
of 1,1 with respect to xl at Po (note that it is tacitly assumed
that separation and 1,1- minimum line are parallel to each other
and only slightly curved [ 92].):

(12.10)
, 13

Now, according to the assumptions made, ,13 = 0 at Po' but not ,23
in general. Furthermore, in general, ,~~ is finite and not equal to
zero at separation, except for the flow in a plane of symmetry
x l = constant(then the flow is directed completely in x 2 -direction).
Separation occurs only if ,23 O. In three-dimensional flow this
indicates - together with ,13 0 - singular separation points
which exist e.g. in lee- and windward symmetry planes of a body at
angle of attack. Thus the derivative 1,1'1 does not vanish at Po'
in general.

In order to determine the location of the obviously present 1,1-


minimum line the, series expansion (12.9) is used in the expres-
sion (12.10). To alleviate the computation compressibility is being
neglected, and so are the derivatives of the transport properties.
The first derivative in x 1 -direction at the pOint P (see Fig. 12.4
m
a» reads:

(12.11)

13 1
("1 13 2) 13 + (~23 23 1 23 2) 23
x + "2 x T'1 ' + T'1 x + T'2 x T'1
P
o

A minimum of 1,1 occurs if 1,1'1 0, and if further 1,1'11 > 0


in P
m

- 164 -
( T 13)2 + T 13 13 23 2 23 23
(12.12) I T I , 11 Ipm ,1 T , 11 + (T, 1 )
IT I
+ T T , 11
Ipm -
2
Ipm
( ITI , 1)
IT I

With ITI'l = 0 at Pm by definition, the shear stress components or


their second derivatives are required to be sufficiently small (if
negative at all), in order to render ITI'll positive, which usually
is the case. From eq. (12.11) the location of the ITI- minimum line
follows:

13 2 23 2
(T'1) + (T'l)
(12.13)

23 23
T T , 1
13 13
T , 1 T'2 +

Because of the linearisations made the ITI- minimum line is locally


approximated by a straight line. The distance between separation
and ITI- minimum line is obtairied from eq. (12.13) for x; = 0:
2323
T T'l
(12.14 )
13 2
(T'l) +

and the inclination of the ITI- minimum line with respect to the
xl-axis is:

13 2 23 2
(T'l) + (T'l)
(12.15) tg 1/I m Ix2 o p
o

Note that in the case of separation in both two-dimensional and


plane-of-symmetry flow in xl-direction, T23 = 0, while T!~ # 0, and
1
therefore xm = O. Thus separation point and ITI- minimum coincide

- 165 -
in these cases, according to eq. (12.14). Because the inclination
of a skin-friction line at P reads:
m

23 23 1 23 x2
(12.16 ) tan 1jJw
• +"1 x +"2
13 1 13 x2
T , 1 x + T'2

it is evident that the 1.1- minimum line at P


m
is not a skin-fric-
tion line. The reader may easily verify the following conclusions
which result from eqs. (12.14) and (12.15):

1. x~ is positive or negative depending on the signs of .23 and


T~~ , and the sign does not depend on the curvature of the sepa-
ration line (see next pages).

2. In the case of separation at an infinite swept cylinder, with


the x 2 -axis parallel to the generator and ~2 = 0, x 1 remains
a x m
finite, and 1jJm becomes n/2. Hence separation and ITI- minimum
line are parallel to each other and to the generators of the
cylinder.

Since it is difficult to determine the separation line experimen-


tally, while the 1.1- minimum line can be obtained in experiment,
see e.g. [ 51, 99], it is useful to have a relation for the distance
between both lines based on data on the 1.1- minimum line. Such a
relation can be obtained from the series expansion (12.9) for the
shear stress on the separation line, assuming again that separation
and I.j-
minimum line are approximately parallel to each other, and
with the x 1 -axis cOinciding now with the minimum line: 1.1-

(12.17) -
T
13
---,-:3
T , 1
I
m

- 166 -
The curvature of the separation and of any skin-friction line is
defined by (.a: local Cartesian coordinates) :

(12.18) Ww,s

where s is measured along it. With

a ax a a a a
(12.19 )
as as ax a
= cos 1/Iw
;;r + sin 1/I w
~
=

13 23
• ;;r
a •
+ --
a
I• I I• I ~

equation (12.18) yields

( 13) 2 23 + 13 23 13 _ (~23)2 13
(12.20)
• "1" - "1) , "2

13
Therefore the curvature of the separation line at Po is with • 0:

13
"2
(12.21)
-~ P
o

In the special case of the flow past an infinite swept cylinder


the curvature of the separation line becomes zero which was found
earlier in the discussion of eq. (12.15).

A closer inspection of the pattern of the skin-friction lines in


Fig. 12.4 reveals an interesting property. In the backward-flow
region the skin-friction lines are negatively curved, including the
separation line. In the forward-flow region they are positively
curved, except in a very narrow region near the separation line
with positive curvature, which is asymptotically followed. Thus the
sign of curvature changes somewhere in that narrow region. There-

- 167 -
fore each skin-friction line in the forward-flow region must ex-
hibit a point of inflexion near the separation line. If the sepa-
ration line is positively curved, the points of inflexion lie in
the backward-flow region.

Introducing the series expansion (12.9) into the relation (12.20)


for the curvature and requiring W to vanish yields an equation
w,s
for the location of the point of inflexion on each skin-friction
line, forming a points-of-inflexion line:

The coefficients a. are functions of the Cartesian wall shear stress


~

components and derivatives thereof at Po:

23 13 23 13
(12.23) a1 1: , 1 h'1 1:'2 - 1:'2 1: 23
, 1) }
P
0

23 13 23 13 13
(12.24) a2 1: [1: , 1 h'2 - 1:, 1) - 2 1:'2 1: 23
, 1] }
P
0

(12.25) a3 - 13
{ 1:'2 13 23 13 23 }
(t'1 1:'2 - 1:'2 1: , 1 )
Po

(12.26) a4 - { 1: 23 1:'2
13 23
(1:'2 + 1: 13
, 1) }
P
0

(12.27)

(12.28) 13
1:'2

According to eq. (12.22) two points-of-inflexion lines are present.


One of the lines is very close to the separation line while the
other is far away as has been found in numerical studies in [ 921.

- 168 -
For the special case of separation at an infinite swept cylinder
only one solution remains:

(12.29) - { 13)2 23
( T'l o
T

Thus in this case the points-of-inflexion line coincides with the


. 13 23
separation line, S1nce both T'l and T are assumed to be finite.

The solutions of eq. (12.22) for x 2 0, that is for the distances


between the points-of-inflexion lines on the xl-axis and the sepa-
ration line, reads:

(12.30) 1
x ip 1,2 ={ 23
2 T'l 13 23
(T'l T'2 - 13 23) }-1
T'2 T'l
Po

.{ T
23 13 13 23 13 23
[T'1 (T'l - T, 2) + 2 T'2 T, 1 ] :!:

23 13 13 23)2 + 4 13 23]1/2}
:!: T T'l [IT'l T'2 T'2 T'l
P
0

At least one of the solutions of eq. (12.30) is small compared


with the characteristic length of the problem, see [ 92]. When in
boundary-layer computations a break-down of the solution occurs it
can usually not be decided whether actually a separation line is
encountered, or whether simply the computation scheme is wrongly
oriented. In both cases the domain-of-dependence condition is vio-
lated, see chapter 6. The investigation of both experimental and
computed boundary-layer data in [ 92], however, has shown that the
ITI- minimum line and the pOints-of-inflexion line are always quite
close to the separation line.

Thus, although the separation line cannot be located exactly, a se-


paration criterion can be formulated: separation occurs, that is,
the line is in the immediate neighbourhood of the pOint of break-
-down, if

*1 *1
(12.31) ~xm « L and ~xip « L

- 169 -
*1
where 6x denotes the physical length increment and where L is the
characteristic body length.

Fig. 12.5 represents a sample result from [ 92]. Po is one point of


break-down of the prediction of the transonic boundary-layer on a
finite swept wing [ 43].

Swept wing s:
~2
= 0.60833

suction side = 0.82003



turbulent tJ' = 1,.31,2"
I'
101",= 0.8i Re= 1.09' 10 7 ,aXA = 3.27·10 -4
I'
'PLE =27.5; 0/ = I· .o1xm =-0.00133
l' -6
W'-' 3.385 .o1x1p 112= 8.5·10 '13.037
A,J

I.. I "ft~
me f-POin,' of infl~xion Ii~e
I~I-mmlmum
x"~
I I I
I--'inal separation line

0.002
provisional separation line ~XA
I'

I -,ax~

·x"
• -I
- '1'." -
1
1 Po Ps
a
r-t--. I'
2 r2
t1x;p, I e"
X,5

-0.002
t1x~" ~J
.o1X:p, I
~-'
:.I,J ~

-0.002 0.002

Fig. 12.5 ITI- minimum-, points-of-inflexion- and separation lines


from a sample calculation in [ 92].

Due to the lack of detailed information the direction normal to the


main-marching (chordwise, ~1) direction was taken as provisional
separation line (xa-system). The ITI- minimum as well as the points-
-of-inflexion line were then determined. These lines were approxi-
mately parallel to each other but not to the provisional separation

- 170 -
line. Taking recourse to the assumption of parallelism of all lines,
aI
a new separation line was constructed (x -system). Thus a final
approximate location of the separation line can be found via eg.
12.17. Note the very small distance between the lines. Fig. 12.6
shows the characteristic pattern of the skin-friction vectors in the

, 2
5wt'pt wing So =0..60.8; So =D.82j e = 0"
"'",,= 0.8; Rt' 7.0.9· 70 1 M~ =-0.0.0.70.7
~LE' 27.5"; ex = ,. .I1x! = 0.000.32/3.93
'P1!2

A,s =0..0.55'I ITI",=3'1O'


suction sidt', turbu/t'nt ~-J

-0.01 -0.005 0.005 0.01

Fig. 12.6 Distribution of the skin-friction vectors in the neigh-


bourhood of the break-down point corresponding to the
flow case of Fig. 12.5. ge is the direction of the ex-
ternal inviscid flow at Po'

neighbourhood of Po' For the sake of clarity only the location of


the 1,1- minimum line and of the points-of-inflexion line on the
xl-axis are indicated. The provisional separation line is shown as
an arc of the circle, see also Fig. 12.5, with its radius of curva-
ture obtained from the inverse of ~ (eg. 12.21).
w, s

- 171 -
12.3 Local Topology of Attachment Lines

The converse of separation lines are attachment lines which divide


diverging forward and backward flow. As was already stated, both,
the inviscid external streamlines,and the limiting wall streamlines
show this behaviour, which is in contrast to the situation near se-
paration lines. In the following the attachment-line flow near the
leading edge of wings or related bodies is considered: first the in-
viscid external then the limiting wall flow. For details see also
SCHWAMBORN [ 781.

12.3.1 External Inviscid Flow

In general the attachment line of the inviscid flow near the leading
edge of a finite wing, in particular at an angle of attack, is
slightly curved, see Fig. 12.7. Here the inviscid flow, whose

ONERA M6 Wing
ex =15°
x 2'

x 2'

(ower upper
surface surface

3'
x
" X

Fig. 12.7 The attachment line at the leading edge of the finite
swept and tapered ONERA M6 wing (from [ 781).

- 172 -
streamlines are shown, was predicted by means of a panel method.
The streamlines are obtained by integrating backwards against the
flow direction, starting from points far downstream. This approach
was chosen since it is difficult to pick the streamlines in the
neighbourhood of the attachment line such that a clear spanwise dis-
tribution of streamlines is achieved. The pattern of the stream-
lines near the leading edge attachment line, and the fact that the
line is, in general, only slightly curved, suggests an investi-
gation like that for the separation line. In addition, there is a
line like the 1.1- minimum line, where the pressure attains a re-
lative maximum close to the leading-edge attachment line. For the
special case of the non-lifting flow past a symmetrical wing both
lines coincide.

Again, for the sake of simplicity, the flow is studied in a local


surface-oriented Cartesian coordinate system like that in Fig. 12.1.
The surface-tangential velocity components are expanded around Po:

V a + va, e x e + ••• } a,e 1,2 .


P
o

In Po the velocity component v~ normal to the attachment line, a


streamline, vanishes, and since o v 3 = 0 at the surface of the wing,
the momentum equation for the xl-direction reduces to

1 2 3
(12.33) p P'l (x ,x,x - 0) ,

if the convective terms at (x a , x 3 = 0) are approximated via eq.


(12.32) by those at Po. Note that the momentum equation in cylin-
drical coordinates for the circumferential direction yields the
same result.

There is a line at which P'l is zero, and which is assumed to be


approximately parallel to the attachment line:

- 173 -
1 2
(v'l) + 2
2 v
(12.34) x
z 1 1 1
v'2 (v'l v'l
p p
o o

with the distance xl on the xl-axis (for x 2 0):


z

1
(12.35) x
z 1 2 1 2
(v'l) + v'2 v'l
p
o

and the inclination

1 2 2
(v'l) + v'2 v'l
(12.36) tg ljJz 1 1
v'2 (v'l
p
o

Thus, in general, the pressure maximum does not occur at the attach-
ment line. Assuming irrotational flow with v~2 = v~l' eqs. (12.34)
to (12.36) can be cast into the same form as eqs. (12.13) to (12.15).
For the special case of the flow past an infinite swept cylinder
with --"- =0 the location of the pressure maximum is given by the
dX 2
attachment line, according to eq. (12.35), and the inclination be-
comes ljJz = n/2, eq. (12.36). The pressure gradient for the x 2 -di-
rection vanishes, and the quasi-two-dimensional flow, discussed in
chapter 5.2, occurs.

On the xl-axis the second derivative of the pressure p with respect


to xl becomes after differentiation of eq. (12.33)

(12.37) P'll z - p
1 2
(v'l) Ip
o

This relation proves that the pressure extremum is a maximum, as


was expected. The maximum is a relative maximum, since the absolute
maximum is attained at the nodal stagnation point.

- 174 -
The curvature of the attachment line in Po is, similar to that of
the separation line:

(12.38) 1/Iz, s IP P
o o

For the special case of an infinite swept cylinder the curvature


vanishes, of course.

The inspection of eqs. (12.35) and (12.38) shows, because usually


1 2 1 2
(v'1) »v'2 v'1 at attachment lines, that the relative pressure
maximum lies on the convex side of the attachment line. The fact
that apparently streamlines diverge from the attachment line and
not from the maximum-pressure line cannot be explained in the frame
of the present investigation.

A pOints-of-inflexion line exists near the attachment line as in


the case of the separation line, namely on the convex side of the
attachment line. The points-of-inflexion line is governed by the
same eqs. (12.22) to (12.30), however with the wall shear-stress
components and their derivatives replaced by the velocity compo-
nents and their derivatives at Po.

12.3.2 Viscous Flow (Boundary Layer)

The pattern of skin-friction lines on the surface of a wing near


the leading edge is in accordance with that of the inviscid stream-
lines near the inviscid attachment line. As long as the inviscid
attachment line is only slightly curved, the distance between vis-
cous and inviscid lines is small. At the stagnation point both lines
touch each other, and for the special cases of the flow past an in-
finite swept cylinder or symmetry-plane flow (symmetrical cross sec-
tion at zero angle of attack) both lines coincide completely, and a

- 175 -
quasi-two-dimensional solution of the boundary-layer equations can
be derived (see chapter 4 and 5). If the inviscid attachment line
is curved strongly and deviates from a geodesic line [ 78], the
limiting wall attachment line will be curved even stronger, and in
the same sense. The reason is the pressure gradient normal to the
inviscid attachment line which is fully transmitted throughout the
boundary layer. In regions where the external streamline exhibits a
point of inflexion, the flow within the boundary-layer follows with
a certain lag which results in the well-known s-shaped cross-flow
profiles.

The skin-friction lines near the limiting wall attachment line are
governed by the same laws as those near the separation line although
this flow is diverging and the other converging. 1.1- minimum line,
curvature and the points-of-inflexion line can be determined from
the equations given in chapter 12.2. Note again that the curva-
ture of the limiting wall attachment line will usually be of the
same kind as that of the external attachment line because sign
(.~~) = sign (V~2)' in general.

Fig. 12.8 shows the attachment lines and a selected number of exter-
nal and limiting wall streamlines for the case of a very strongly
flattened triaxial ellipsoid at large angle of attack resembling a
wing with varying leading edge sweep (from SCHWAMBORN [ 78]). Be-
cause of the scale of the drawing the small distance between invis-
cid-flow and limiting wall attachment lines is not distinguishable.
Because of the positive angle of attack both lines are located on
the lower surface. Again, in order to produce a clear picture of
the streamlines diverging from the attachment lines, points on the
surface were picked some distance downstream on the lower surface,
and the streamlines were integrated backwards towards the attach-
ment lines. Thus inviscid and limiting wall streamlines can be com-
pared which pass through the same point downstream. The integra-
tion of the streamlines of the upper-surface flow started arbitrari-
ly at the leading edge, but at the same spanwise station. For the
span station far outboard two starting points for the streamline
integration were chosen for the lower-surface flow to show that
the difference between viscous and inviscid streamlines decreases

- 176 -
the closer to the attachment lines the common point is chosen in
spite of the strong three-dimensionality of the flow. Note the re-
duction of the distance between the inviscid and limiting wall

2' Effipsoid 3: 1: 0.125 2'


X X
a= 15°

attachment
line
B
leading
edge
---
upper
surface
-----
-----
limiting wall} "
" ""d streamlmes on lower
mVlscl surface

starting pOint for backward


streamline integration

3'
stagnation point 5 x" A X

Fig. 12.8 Streamline pattern near the attachment line on a tri-


axial ellipsoid (3 : 1 : 0,125) at an angle of attack of
15 0 in incompressible flow [ 78].

streamlines close to the quasi-two-dimensional flow regime in the


2'
symmetry plane x = O. It is also mentioned that, strictly spea-
king, first-order boundary-layer theory, which was used to predict

- 177 -
oj
30r 11111 c)
T:* 13 ~..;( I IIIII I bJ
X' :conS f. T:* 23 /T:* I
8
-...J
co "1
Il"·_M 24f- /I I I f I I I I )(7: const.
20~ 1111111 VI

.l:con st

mag nitu de
and b) ,'2 3). and of the
sica l com pon ents ( a) ,"3
Fig . 12. 9 Dis trib uti on of the phY l elli psO id
(c) iT* I) nea r the lea din g edg e of the tria xia
of the wal l she ar str ess :;-*(13 IRe ref / Pre f(V ref} 2}.
Fig . 12. 8, [78 J fT* a3 =
the flow, looses its validity if the very tip region is approached
since the nose radius becomes smaller and smaller (see chapter 9),
while the boundary-layer thickens more and more.

Fig. 12.9 shows and compares the distributions of the wall shear
stress components and of the total wall shear stress itself in the
neighbourhood of the attachment line depicted in Fig. 12.8. The
distributions are given versus the chordwise (xl) surface parameter
for several spanwise stations (x 2 ). To indicate the variation in
spanwise direction x 2 , the values along selected lines xl = const.
are presented as solid lines. In addition the shear stress along
the attachment line, see Fig. 12.8, is displayed. Note the large
variation of the wall shear in xl-direction towards the leading
edge (+ xl) in comparison to that in negative x 1 -dir-ection towards
the trailing edge on the lower surface. The h*l-plot shows clearly
the IT I -minimum line, and also that there is apparently a differ-
ence in location between the IT I -minimum line and the attachment
line, although they are quite close to each other, see [ 92].

Note, finally, that SCHWAMBORN [78] derives the flow picture,


presented in chapter 9.2, also by discussing the surface-tangen-
tial components of the vector equation y x grad p = o.

- 179 -
13. Similarity-Type Transformation of Boundary-Layer Equations
in Contravariant Form

13.1 General Remarks

In particular for three-dimensional boundary layers it is worth-


while to consider transformations of the boundary-layer equations
which keep the thickness of the boundary layer in the computational
space nearly constant. In two-dimensional laminar boundary-layer
predictions similarity-type, such as Blasius or Levy-Lees, trans-
formations generally achieve the desired goal. In turbulent bound-
ary layers the advantage of such transformations is not entirely
clear since the Blasius or Levy-Lees scaling had been devised for
laminar flows, and turbulent boundary layers grow at a faster
rate. One of the main features of such transformations is that
the flow variables are scaled by the local values at the "edge"
of the boundary layer which yields a constant value of one there,
and which prevents in particular the transformed velocity profiles
to change drastically in the streamwise direction, at least in the
absence of a sustained strong adverse pressure gradient. Such
transformations have been maintained for three-dimensional prob-
lems (see e.g. BLOTTNER and ELLIS [15] or CEBECI et al. [18])
although it is, in general, not possible to scale the crosswise
velocity component with its edge value since that one may change
its sign.

NASH and SCRUGGS [66,67] do not scale the dependent variables with
the corresponding values at the "edge" of the boundary layer.
Instead they use apprdpriate constant reference quantities (see
chapter 2). The scaling of the boundary-layer thickening is
achieved by simply adopting suitably varying distributions of a
fixed number of node points across the boundary layer. For laminar
flows a uniform distribution is chosen with

(13 • 1 )
-3
x ~ 0 m/M

in the physical domain, where ~ z 1.25 and 0 is taken as the dis-


tance away from the surface where 99.5% of the external velocity
is reached. M is the total number of increments bX 3 = bx 3 /L
re f

- 180 -
across the layer of thickness n &,
while m denotes the local dis-
tance in the computational domain x 3 = m ox 3 . For turbulent flows
a non-uniform distribution is used which provides an increased
density of the mesh points in the region of large gradients close
to the surface. According to the various regions of the turbulent
boundary layer - viscous sublayer, logarithmic region and wake
region - the expression in terms of physical quantities is

(1 3.2)
-3
x (1
<I>
+
&
<p) M [c 1 (m - M) + n Ml ,

where <p = <p (m) is such that <p becomes large as m ~ M. Thus in
the outer part of the boundary layer a uniform distribution of
mesh points is obtained. In the inner part equation (13.2) is
controlled by

(1 3.3)

Thus the choice of the constants c 1 to c 4 , where c 2 determines


the distance from the surface at which the first mesh point is to
be placed, and where c 3 is taken to be proportional to the additive
constant in the law of the wall, determines the distribution of
M node points in turbulent flows. For the example of incompressible
two-dimensional Blasius and turbulent flat-plate flows [67] it
is shown that 30 points across the boundary layer givg excellent
accuracy. It is, however, not clear how a smooth transition with
respect to the node-point distribution is made, when an initially
laminar flow becomes turbulent. The transformations (13.1) and
(13.2) represent effectively a scaling of the distance normal to
the surface by means of the boundary-layer thickness, which is
approximated as a finite distance in boundary-layer computations.
First derivatives of &with respect to the surface-tangential
directions appear in the boundary-layer equations, and have to be
discretized as well.

A scaling of the surface-normal coordinate similar to that used


by NASH and SCRUGGS is employed by LINDHOUT et al. [581:
~ = f (x 3 , 6). For laminar flows f is chosen to be simply the
ratio x3/~:

- 181 -
(13.4) r. x-3 /6-

while, for turbulent flows, the normal coordinate is stretched


near the wall. Note that here 6 is the distance away from the
surface of the body at which 95% of the inviscid velocity are
reached [58). Near the wall the transformation is independent
of 6
and constant throughout the complete computation. A preas-
signed value of f'w = df(6;O)/dx 3 is incorporated which depends
-3 -
on the Reynolds number. For the interval 0 ~ x ~ 6 the average
value of f' is chosen to be 1/6, the same as for laminar flow.
For large x 3 the transformation tends to be linear, and f' (0,00)
is taken to be some preassigned fraction r of 1/6: f' (0,00) = rio.
The following relation satisfies the above requirements:

f'
w + a x3
(13.5)

where x3 -3 -
x /6, a br/6 and b (0 f'w - 1)/(1 - r).

The transformations (13.4) and (13.5) allow non-equidistant wall-


normal meshes in the computational domain if wanted, contrary to
equation (13.2). A smooth transition from laminar to turbulent
flow with respect to the transformations seems again not feasible
other than by interpolation for the "turbulent" mesh, or by using
a weighted average of both transformations by means of which the
turbulent mesh can be turned on gradually while the laminar mesh
is disappearing. LINDHOUT et al. state that C has to be chosen
-3 _ max
such that x max/6 ~ 2, to obtain sufficient accuracy. Note, that
all three physical velocity components are made dimensionless with
the help of the inviscid velocity component in one direction. Com-
pressibility effects are assumed to be approximated with the help
of the Crocco relation in terms of density and velocities, and
the density is referenced with the corresponding external value
as well.

Scaling with another length is suggested by McLEAN [61,62). The


normalizing length for laminar as well as turbulent flows has
been chosen to be 6~, the three-dimensional outer length scale in

- 182 -
turbulent flows:

(13.6)

where u; represents the inviscid external velocity, such that the


transformation reads:

(13.7)

For turbulent flows, it is reported that it proved to be sufficient


to terminate the grid and to apply the outer flow boundary con-
ditions at about ~max = 15. Note that McLEAN employs a non-equi-
distant mesh with geometric progression where the constant ratio
of adjacent steps K = 6~k/6~k_1 is typically 1.25 with 6~ • 0.001
for full-scale turbulent flows, resulting in about 37 pOints across
the boundary layer [62], if ~max = 15. For laminar flows a smaller
value of ~max is sufficient [62]. The dependent variables are
chosen to be some kind of defect quantities, e.g. f' = 1 - pv*1/
(PeU;) for the x 1 -momentum equation. The temperature is normalized
with the corresponding inviscid value. It is interesting to learn
that McLEAN uses essentially the form of the governing partial
differential equations as given by CEBECI, see chapter 13.2, i.e.
with terms involving third derivatives f'" and quasi-linear pro-
ducts such as ff" '. Nevertheless he uses only three points to
approximate derivatives, since he locally linearizes the coeffi-
cients involving f (and g, see chapter 13.2), and is thus able to
introduce a new variable F = f' which reduces the highest order
of derivatives to two.

In order to be able to evaluate the usefulness of the similarity-


type transformation, as is suggested by BLOTTNER [14,15] or CEBECI
et al. [18], such a transformation is being applied here to the
governing equations in the contravariant form as is given in the
main chapters. It is noted again that the similarity-type trans-
formation scales the flow, in general, only in one direction. Here,
without restriction, the x 1 -direction is chosen which might coin-
cide with the chordwise direction for the flow past a wing or with

- 183 -
the direction of the major axis for the flow about a body at a
not too large angle of attack. If external streamline coordinates
were used, one would choose xl to coincide with the external stream-
lines. Note, however, that the definition of external-streamline
coordinates is, in general, a difficult task in particular if
the inviscid velocity components result from a numerical predic-
tion. Note also that if the similarity transformation which was
originally devised for laminar flows is used in turbulent-flow cal-
culations, it is necessary to use a non-equidistant wall-normal mesh
with geometric progression to cope with the growth of the boundary-
layer thickness. CEBECI et al. report a maximum number of about
30 points across the boundary layer for the incompressible flow
about wings at moderate Reynolds number

13.2 General Three-Dimensional Flow

Consider the boundary-layer equations (2.1) to (2.4) and the


relations (3.4) to (3.8) , associated with the mass-flow displace-
ment due to the boundary layer, for steady-state flows.

Introduce the following Ansatz:

1
(13.8) kOl P v ""3
2
kOl P v .", 3

0 = T/Te ,

where the subscript 3 means partial derivative with respect to x 3


which is related to a new similarity wall-normal coordinate ~ by

(13.9) d~

This relation can be rewritten as follows:

dl; (Re ) 1/2 L


e,x 1 Pe

which effectively represents a scaling of the surface-normal coor-


dinate x 3 with the square-root of the local Reynolds number

- 184 -
Re e ,x 1 = Pe v~ xl/~e' with the chosen main coordinate xl and with
the density ratio piPe' Note that, since constant reference
quantities are used for all variables (see chapter 2), the usual
boundary-layer stretching with (Re ref ) 1/2 of the coordinate x 3
evolves naturally, and that Re e,x 1 is actually referenced with
Re ref . It is assumed that non-dimensionless quantities are
used, so that the surface metric tensor coordinates and
therefore the metric factors are non-dimensional as well. The
functions W and ~ are assumed to have the form

(13.10)

Q = 1,2,

where v; denotes a reference velocity for the velocity component


v 2 which does not need to be a constant, and where

(13.11 )

Thus the following relations result from equations (13.8):

13.12) f' v1 u
v e1

g' v2 v
2"
vr

where the prime denotes partial differentiation with respect to


the coordinate ~, and where f' and g' are abbreviated by u and v,
indicating that those quantities replace velocities in the momen-
tum equations. This is done because the above derivatives them-
selves will later be treated as new dependent variables.

There are two well-known ways to use the similarity-type trans-


formation. One way is to use the integrated continuity equation
(2.1)

(13.13 )

- 185 -
with

(13.14) ljJ , 1

(p " v 1 x 1 )1/2 f'1


e "'e e

<1>'2

c
r

in order to replace the convective term in the x 3 -direction. There-


by the continuity equation need not be integrated explicitly. How-
ever, third derivatives need to be approximated numerically.
CEBECI et al. [ 18] split any higher-order equation into first-
order equations, in order to be able to apply their Keller-Box
scheme, and thus no difficulty occurs for the numerical approxi-
mation. Other schemes, however, need more points than usual in
their finite-difference molecule to assure second-order accuracy.
That is why BLOTTNER [ 14] choses the second way. He integrates
the continuity equation explicitly where the present formulation
yields the following:

(13.15) (K 01 u), 1 + (K 02 v), 2 + (K 03 w), r, o

where

(13.16 ) w = ~'1
V e1 U + ~'2
v2
r v + ~'3 v
3

is the new surface-normal velocity component. Note that the easiest


way to obtain eq. (13.15) is to make use of VIVIAND's results on
transforming equations in conservative form [132]. The new sur-
face-normal velocity component w is, as well as the surface-normal
coordinate x 3 , stretched with (Re f) 1/2, which follows from
re
equations (13.9) and (13.13). In equation (13.15) ~ denotes the
new transformed x 3 -coordinate, equation (13.9). The coefficients
KOn now contain the metric factor k 01 ' which is the square root of
the determinant of the surface metric tensor, and properties of
the external inviscid flow field:

- 186 -
(13.17)

Since, as usual, only first and second derivatives of the new va-
riables are to be discretised, this approach is taken here.

Introducing the relations (13.8) to (13.12) into the governing


boundary-layer equations results in the following equations in non-
dimensional form with the continuity equation given above.
Momentum equation for x 1 -direction

(13.18)

where the coefficients K1n are the following abbreviations:

(13.19) K11 x 1, K12 c x1, K13 (k 11 + v e1 '1/ v e)


1
x
1
r

1 1 1 1
K14 c
r
(k 12 + v e '2/ v e) x , K15 k13 (C r )2 x ,

K16 x1/v~ K17 Pe [K 13 + K14 v e + K15 (ve ) 2)

Momentum equation for x 2 -direction

(13.20)

with

- 187 -
1 1
(13.21) K21 x Kll K22 C x K12
I
r

1 2 2 1
K23 (k 21 /c r ) x K24 = (k 22 + v r'1 /v r ) x

2 2 1
K25 c (k 23 + v r'2 /v r ) x K26 x 1 /v 1 K16
r e

K27 Pe [K 21 v e '1 + K22 v e v e ' 2 + K23 + K24 v e +

Energy equation

(13.22)

V,
r,

with

1 1
(13.23) K41 (Te ' 1/Te) x K42 x K11

1 1
K43 cr (Te ' 2/Te) x K44 c x K12
r

-1
K45 X 1 /v 1 K16 K46 (Il e Pel K18
e I

K47 x 1 /T K48 c x 1 /T
e r e

1 2
K49 k41 (v e ) /(Il e P e Te)

- 188 -
Note that the quantities at the "edge" of the boundary layer are
referenced with the constant reference quantities defined in chapter
2. If CEBECI's approach would be followed equation (13.13) would be
needed in the form :

(13.24)

and the governing equations would change accordingly.

A remark is appropriate here concerning the choice of v;. The


optimal choice v 2 = v 2 is, in general, impossible since v 2 may
2 r e 2 1 e
vanish while v ; O. Choosing vr = ve would simplify equations
(13.15) to (13.23) appreciably but, in certain cases such as close
to the leading-edge attachment line on finite wings, v~ as re-
ference quantity can result in rather large variations of v in the
downstream direction since v~ is very small there. Therefore, often
2
a constant reference velocity component vr is used.

Equation (3.5) for the mass-flow displacement thickness 01 remains


nearly unchanged

(13.25)

- 189 -
but the two-dimensional displacement thicknesses change as follows:

1
I,( x 1/2 0 P e
(_e_ _ )
(13.26) 01 1 1 I (P- - u) d~
x Pe v e 0

1
I,( x 1/2 0 P e
(_e_ _ )
(13.27) 01 2 1 I ( - - ~)
v
d~
x P e ve 0 P e

The equivalent inviscid wall-outflow velocity is obtained from


equation (3.8) by making the same changes as in equation (13.25).
The equations for the other displacement thicknesses undergo simi-
lar changes.

The external and wall boundary conditions needed are according to


those given in chapter 2, and are not discussed here. The reformu-
lation of the boundary conditions for the following paragraphs
will be left to the reader as well.

13.3 Stagnation-Point Flow

In order to arrive at the boundary-Iay~r equations in similarity-


type contravariant form for the flow at the three-dimensional
stagnation pOint, the symmetry properties for the stagnation-point
2 2
flows as discussed in chapter 4 are used again. With v = v and
1 r2 e
the usual assumptions of linear variations of v and v near the
stagnation point (xl 0, x 2 = 0: v 1 = 0 , v 2 _ 0):

(13.28) + ••••• along the coordinate x 2 o

(13.29) + ••••• along the coordinate xl o

one obtains the equations given e.g. by LIBBY [ 53) if the limiting
process for xl ~ 0 and x 2 ~ 0 is applied to the equations derived
in the preceding chapter 13.2. The transformation relation (13.9)
yields then

(13.30) dr; [ v 1 'l /( P I,(e ))1/2 P dx3


e e

- 190 -
and the dependent variables u and v, equations (13.12), change
according to equations (13.28) and (13.29):

(13.31) f' = v,1 1 Iv e1 , 1 AllAl


e
u

(13.32) g' v

2 2 2 2
The notation vr and v r '2 is kept here. Although v r '2 v e '2 is a
good choice at the stagnation point since v 2 '2 = v 2 '2 results in
r 2 e 2
Ve 1, this is not so in the symmetry plane x = 0 where v = 0,
since v!'2 generally changes its sign with increasing xl while v~2
does generally not at the same time. However, one could use there
a constant, e.g. v;'2 = A; = v!'2S = const., where the subscript S
denotes the stagnation point. This choice can result in large va-
riations of V along both, surface normal and the coordinate x 2 = o.
Therefore it would also be appropriate to use a constant value v;
(the same as for the general field calculation), and to consider
~ 2 2
v = v'2/vr as new dependent variable (see equation (13.12». Here
the definition (13.32) with A; = A! as reference quantity is used
since it is then easy to arrive at the usual similarity-type form
of the equations if wanted. The dependent variables are then U, V,
wand B according to equations (13.31), (13.32), (13.16) and (13.8).
Continuity equation

(13.33) o

with

(13.34) V

Momentum equation for xl-direction

(13.35 )

with

- 191 -
(13.36)

Momentum equation for x 2 -direction (note that now v in equa-


e
tions (13.21), since v 2 v2)
r e

(13.37)

where

(13.38)

Energy equation:

(13.39)

where

(13.40)

The equations for the mass-flow displacement thickness and the


equivalent inviscid wall-outflow velocity are those given in chap-
ter 4 with the following definitions for the two-dimensional dis-
placement thicknesses:

lJ. 1/2 <I P


(13.41)
°1x a (_e_) J (-e - H) d~ a = 1,2
Pe
A1
e
o P

where H = U for a = 1 and H = V for a = 2. The usual similarity


form [53 1 of the governing equations is obtained if the wall-
normal mass flux is replaced by the limiting form of the equation
(13.24) :

(13.42)

where U _ f,~ and V _ g,~, see equations (13.31) and (13.32).

- 192 -
13.4 Plane-of-Symmetry Flow

The symmetry plane is assumed to be erected along the coordinate


line x 2 = 0 (see Fig. 5.1) where v 2 = 0, and v 1 # 0 for xl > o.
Therefore the following dependent variables are chosen:

1
(13.43) f' --,
v
v
= u
e

2
2
v, 2 A2
v
g' 2-> -2- V
A2
vr v r '2 r

w follows from equation (13.15), and the transformation rela-


tion for the wall-normal coordinate is given by equation (13.9).
The reference quantity v 2 can be assumed variable as long as
r
v 2 '2 = A2 # 0 along x 2 = O. Note that this condition rules out the
r r 2 1 1 2 1
choice vr ve there, if v e '2 (x = 0) = - ve a 11 '2/(2 all) va-
nishes which can be assumed to happen in general, see chapter 10.
A constant is then introduced as a convenient choice for A2 (another
choice would be to take the maximum value A2 (x 1 ) for A2r (x 1 +
1 max r2
+ 6x ); and according to the discussion in chapter 13.3 A could
2 2 r
be replaced by a constant value for v as well). Here A is re-
2 r 2
ferenced by Ae at the stagnation point: AeS.

With

(13.44)

the following equations are obtained:


Continuity equation (see eq. (13.15))

(13.45) o

where

- 193 -
(13.46)

Note that the abbreviations Kmn are determined by the expressions


2 2 2
given in chapter 13.2 by replacing v by v '2 = A S' and by letting
2 2 2 2 r r e 2 _
v = 0, v = v /v r ~ 0 for x ~ 0 as well. Also note that v r '1 = o.
Momentum equation for x 1-direction (see eq. ( 1 3 • 18) )

2
( 13.47) K11 u u'1 + K13 (u) + K16 w u,~

with

(13.48)

Momentum equation for x 2-direction (see eq. (13.20), and note that
the equation has to be differentiated with respect to x 2 )

(13.49)

where

(13.50)

Energy equation (see eq. (13.22»

(13.51)

In order to predict the displacement effect due to the symmetry-


plane flow, equations (5.16) and (5.19) can be used. The two-di-
1
mensional displacement thickness with respect to the v -velocity

- 194 -
component is given by equation (13.26), that one with respect to
the v 2 -component by equation (13.27) if v and ve are replaced by
V and Ve. The remaining displacement thicknesses follow in a simi-
lar fashion.

13.5 Locally-Infinite-Swept-Wing Concept

Using the concept explained in chapter 5.2.3 yields the following


set of governing equations with the definitions of u, v, wand 0
taken from chapter 13.2, and with c r defined in equation (13.11),
2 2
however vr -F ve.
Continuity equation

(13.52)

where

(13.53)

+ are defined in chapter 5.2.3, while the expres-


Note that all k mn
sions Kmn are given in chapter 13.2.

Momentum equation for xl-direction

(13.54) 2 + 2
Kll U u'l + K13 (u) + K14 U v + K15 (v) +

where

(13.55)

Pe
[ K 13 + K+
14 ve + K 15 (ve ) 2

Momentum equation for x 2-direction

(13.56)

- 195 -
where

(13.57)

Energy equation

(13.58)

The partial differential equations (5.58) and (5.59) to determine


the displacement of the inviscid flow due to the boundary layer
hold with equations (13.26) and (13.27) to calculate the two-dimen-
sional displacement thicknesses.

13.6 Initial Data for the Locally-Infinite-Swept-Wing


Solution

Initial data for the integration of the governing boundary-


layer equations, given in the preceding chapter 13.5, are obtained
by considering those equations term by term for the limiting case
1 1
x + 0 and v + 0 at the same time (see also chapter 5.2.4). The
dependent variables become then

(13.59) f' Al/Al u


e

g' v 2 /vr2 = v

- 196 -
and w and e are defined as usual (see e.g. chapter 13.2). The trans-
formation relation for the wall-normal coordinate is given again by
equation (13.30) assuming a linear variation of v 1 to hold near
1 2 2
x = 0 along x = constant. The reference quantity vr in equation
(13.59) should be consistent with the choice made for the equations
2 2
in chapter 13.5. However, in general, the choice v - v would be
1 r e
possible at x = 0, and would result in ve - 1. Note that in the
derivation of the equations in chapter 5.2.5 as well as in that of
*2
the following equations the assumption of v'1 = 0 has to be made in
order to arrive at a closed set of quasi-one-dimensional equations.
The governing differential equations read as follows:
Continuity equation

(13.60)

where

(13.61)

+ is defined in eq. (13.34).


and where K03

Momentum equation for xl-direction

(13.62) (U)
2 + U v + K+ (v) 2 +
+ K14 15

where

(13.63)

and with K+
16 and K18 given by equations (13.36) and (13.19).

Momentum equation for x 2-direction

(13.64) K+++ 2 +
25 (v) + K26 w V't =

- 197 -
where

(13.65) K++++ K+++ (v )2


27 Pe 25 e

and K+
26 and K28 are given in equations (13.38) and (13.21).

Energy equation

( 13 • 66 )

where

(13.67)

+ and K46 are defined in equations (13.40) and (13.23).


and K45

The displacement thickness 01 and the equivalent inviscid wall-


outflow velocity are obtained from the expressions (S.71) and (5.72)
with the two-dimensional displacement thickness 01 ' a 1,2,
xa
determined by equation (13.41) with H = U for a = 1 and H = v/ve
for a = 2.

- 198 -
Appendix A: Basic Geometrical Relations

In this chapter the most important geometrical relations are put


together, they can be found in textbooks [ 5, 7,12] or in [ 2].
Note that all lenghts are considered as non-dimensionalized with
a reference length L- ref , but that x 3 (and v 3 ) are not yet stretched
with the square root of the reference Reynolds number.

A.1 General Coordinates

The Cartesian right-handed coordinate system xi', as depicted in


Figure 2.1, is used as basic reference coordinate system throughout
this book. The base vectors of the Cartesian system are the unit
0'
vectors eo, in the directions x~ Suppose that a set of curvi-
-~ 0

linear coordinates x J is given in such a way that the coordinates


of the position vector R' (see Fig. A.1)

(A.1) R' = R' (!') =g' (~),

i,j 1,2,3 .

are known (note that x j may be parameters only, and not necessarily
lenghts measured along the coordinate lines). Then the coordinates
8~ of the covariant base vectors g. of the xj-coordinate system
J -J
are determined as follows:

(A.2) a
.... J
0 = -R', J 0

~i'
i,j 1,2,3.

-i'
where Bj are the partial derivatives of the coordinates of the
position vector R'

(A.3)

*)The Einstein summation convention is used.

- 199 -
The covariant metric tensor which is symmetric then follows as

3
(A.4) E j ,k 1,2,3,
i'=l

and the length element ds is defined by

3
(A.5) (dS)2 -i'
j ,k 1,2,3.
i'=1 '\

For the case of general space coordinates x j the geometrical rela-


tions necessary (A.l) are difficult to obtain in analytic form.
Therefore, often - at least in two dimensions - numerical techni-
ques are employed to generate the mesh (see for instance THOMPSON
et al. [105]). If locally monoclinic surface-oriented coordinates
can be used, e.g. for the calculation of thin shear layers the re-
lations (A.l) are easily derived, as long as the surface can be de-
scribed (note however the limitations of such systems as dis-
cussed in the introduction). The geometrical relations for such
systems are presented next.

A.2 Surface-Oriented Locally Monoclinic Coordinates

In Figure A.l a surface-oriented locally monoclinic coordinate


system is depicted. Such a system is characterized by the fact
that the surface of the body defines the coordinate surface
x 3 = 0, which is spanned by lines xl constant and x 2 = constant.
The third coordinate x 3 is rectilinear and erected locally normal
to the surface of the considered body. The angle ~ denotes the
angle between lines xl = constant and x 2 = constant which may be
non-orthogonal, in general. Using the rules of vector addition the
general position vector R' (see Fig. A.l) is split into two parts
one of which varies line;rly with x 3 :

(A.G)
i,j 1,2,3,
a = 1,2,
- 200 -
x ]'

X =const.
2

l'
X

Fig. A.1: Definition of surface-oriented


locally monoclinic coordinates.

where -r' (xu) denotes the position vector of a point P s on the


surface, and where ~3(xu) is the unit vector in x 3 -direction. The
coordinates of this latter vector are obtained either as follows:

i'
(A.7) a
-3 = [~' (xi' ) I , 3 = ~' , i ' x ,3 ~i'

or from the condition that ~3 is normal to the surface, i.e.


~3 ~ ~1 x ~2' and that ~3 . ~3 = 1. Thus, both, E' and ~3 depend
on the surface properties of the body or, what is equivalent, on
the surface coordinates xu.

Therefore, any point P in the space above convex surfaces (for


concave surfaces at least in the neighbourhood of the

- 201 -
surface) is uniquely defined if the corresponding point P s on the
-i'
surface and the coordinates 6 3 are known. The problem of defini-
tion of coordinates is then reduced to the definition of surface
coordinates. Where physically appropriate, the neglection of the
dependence on x 3 further eases the formulation of viscous-flow
problems as has been shown for boundary-layer problems in the main
chapters of this book.

These surface-oriented locally monoclinic coordinates have the


feature -for general cases- that the coordinate lines in surfaces
o < x 3 = const. are skewed in comparison with the corresponding
coordinate lines in the surface of the body x 3 O. This skewing,
however, is negligible if the thickness of the considered layer
is small compared with the smallest principal radius of curvature,
as has been discussed in chapter 9. In special cases no such
skewing is present, and it vanishes identically for two-dimensional
problems. The indicated skewing of coordinates (see equation (7.11»
is not a principal restriction, but may cause some inconveniences
regarding the representation of results since it may require
interpolation or matching procedures.

Since the concept of surface-oriented locally monoclinic coordinates


yields an exact formulation it may be used in the entire computa-
tional domain as required for the solution of the considered
problem. The only possible limitations occur in cases with concave
surfaces because then the x 3 -coordinate lines converge, and cross
over at some distance x 3 away from the body. This distance in-
creases with increasing radius of curvature. Therefore, in first-
order boundary-layer theory this limitation is eliminated by
definition (for other cases see the remarks made in the intro-
duction) .

In the following the properties of surface-oriented locally


monoclinic coordinate systems are presented. Relations are given
a) for general coordinates (see preceding paragraph) where
appropriate, b) for surface-oriented coordinates for points P on
and off the surface of the body (higher-order boundary-layer,

- 202 -
thin-layer Navier-Stokes, general two-dimensional problems),
c) for surface-oriented coordinates for points P s on the surface
of the body (zero- and first-order boundary-layer problems
neglecting terms involving x 3 ). Coordinates a) are referred to as
general coordinates, coordinates b) as surface-oriented coordinates
with g-metric, and coordinates c) are called surface-oriented
with a-metric. The notation "locally monoclinic" is omitted in
the following.

A.3 Covariant Base vectors of Surface-Oriented Coordinate Systems

Consider a point P s on the surface of a body (Fig. A.1) given by


the position vector r' in the reference system

(A.B) r' = r' (x')


-s = r'
-
(x )
-s
i'
(x s ); i 1,2,3, CI 1 ,2,

where x a are the chosen Gaussian surface parameters (see, for


examples, chapter 10). Then the components of the covariant base
vectors a , corresponding to the xa-directions, are obtained by
-a
differentiating r' with respect to x a [ 7,121

(A.9) a r' Q 3' •


-a - 'a ~i' ·~1 ' "a '

a = 1,2.

Here a!' (note the difference between a!' and s!', the latter
either being partial derivatives with respect to general coordinates
(eg. (A.2)), or especially derivatives with respect to off-surface
coordinates with g-metric, eg.(A.1B)) denotes the partial deriva-
tives of the Cartesian reference coordinates with respect to the
Gaussian surface coordinates:

(A. 10)

- 203 -
The third base vector ~3 is, by definition, orthogonal to ~1 and
~2' and a vector of unit length:

(A. 11) ~3 • ~a = 0, ~3' ~3 = 1; a = 1,2.

Therefore, instead of using in some form equation (A.7), ~3 is


often determined by [

(A.12)

This eliminates the need to establish the general relations of


transformation (A.l), and only surface properties have to be
known. The coefficients 6~' follow from the first equations

(A.13) 6i '
3 I;i/I; ,

where

2'
1;1 61 6 32 ' - 6 31 ' 6 22 '

(A.14) 1;2 6 21 ' 6 31 ' - 3'


82
1'
81

2'
1;3
1'
81
2'
82 - 1'
82 8,

and

(A.1S) I; [ (1;1) 2 + (I; ) 2 + (I; ) 2)' /2 = fa


2 3 = k O'

The definition (A.2) of ~i and (A.9) of ~a together with the


relation for a general position vector (A.6) yield the covariant
base vectors for the surface-oriented coordinate system with
g-metric:

(A.16) a = ',2

(A.17) ~3

- 204 -
-if
Therefore the components 8 j read as follows:

-if 8 i ' + x 3 i'


(A.18) 8a a 8 3 ,a

-i f
(A.19) 83 8i ' i' 1,2,3,
3
a 1,2.

1
The relations (A.18) show why the x , x 2 -coordinates get skewed
when one considers different surfaces x 3 = const. at the same
location (x 1 ,x 2 ): the cosine of the corresponding angle is given
-1/2
by cos ~ = g12 (g11 g22) , which will in general be only con-
stant for constant surface parameters xa'if ~3 = ~3 does not
depend on x a (see also page 73). This obviously holds always in
the case of flat surfaces.

A.4 Transformation of a Vector

The transformation of the contravariant coordinates of a vector


j if
~ from the x -system into the x -system is given by [ 2, 5, 7,12]

(A.20) i,j 1,2,3,

where the coefficients B~' are obtained from equations either


J
(A.2) or (A.18) and (A.19). Note that equation (A.20) represents
the contravariant transformation law for vectors in the generally
used physical form only if the diagonal coordinates of the metric
tensor are identical unity: g(ii) = 1.

If

-1 ' -1 ~ -1 '
81 82 83

(A.21) -2' -2' -2'


B 81 82 83

-3' -3' -3 '


81 82 83

- 205 -
is the transformation matrix for the transformation of the
contravariant vector coordinates eq. (A.20), one finds by inver-
sion the matrix

-,13, , -,13 2 , -,13 3 ,


(A.22) B-' -2
13, ,
-2
13 2 ,
-2
13 3 ,

-3 -3 -3
13, , 13 2 , 13 3 ,

-2' -3' -3' -2' -3' -3' -1' -2' -2' -1 '
(13 2 13 3 - 13 2 13 3 ) (13 2 -"
13 3 - -1'
13 2 13 3 ) (13 2 13 3 - 13 2 13 3 )

1 (13-3' -2' -2' -3' -1' -3' -3' -1' -2' -1' -1' -2'
TBT 1 13 3 - 13 1 13 3 ) (13 1 13 3 - 13 1 13 3 ) (13, 13 3 - 13, 13 3 )

-2' -3' -3' -2'


( 13, 13 2 - 13,
-3'
13 2 ) (13, -" -"
13 2 - 13, -3' -"
13 2 ) (13,
-2' -2'
13 2 - 13, -"
13 2 )

with

(A.23) IBI -,'


13, (13-2'
-3'
2 13 3 - -2' -3'
13 3 13 2 ) +

-, ,
+ 13 2
-2' -3'
(13 3 13 1 - -2' -3'
13 1 13 3 ) +

-1 ' -2' -3' -2' -3'


+ 13 3 ( 13 1 13 2 - 13 2 13 1 ) /g

The matrix (A.22) is the transformation matrix for the inverse


transformation, the explicit equations of which are, with other
techniques, in general hard to obtain:

(A.24) i,j 1,2,3.

Note that IB- 1 1 lat, I is generally called Jacobian of the trans-


formation.
- 206 -
The relations (A.20) to (A.24) hold for all three kinds of
coordinate systems defined at the end of paragraph A.2. In the
case of surface-oriented coordinates with a-metric the barred
transformation coefficients have simply to be replaced by those
without bars defined in equations (A.10) and (A.13).

A.S Partial Derivatives

Between the first derivatives of a scalar function w in the


i' j
x -system and in the x -system the following relations exist

(A.2S) i,j 1 ,2,3

and likewise

(A.26)
i,j 1,2,3.

Higher derivatives are derived by repeated application of the


first derivative, for example:

(A. 27) (w'1') '1'


w'1 ' 1 '

j,k 1,2,3.

In the case of vectors one has to proceed in the same manner


taking into consideration, in addition, the variation of the base
vectors for example in the contravariant representation ~ = vk~k
(see also Appendix B) .

As in the preceding paragraph the relations hold for all three


considered types of coordinate systems, in the case of the use
of the a-metric the overbars on -s
at have to be omitted.

- 207 -
A.6 The Covariant Metric Tensor for Surface-Oriented Coordinates

The relations for the covariant metric tensor g .. for general


1)
coordinates were already given in equations (A.4). For surface-
oriented locally monoclinic coordinates the symmetric metric
tensor takes the form

(A.28)

Since in all equations given for surface-oriented coordinates


the special form of the tensor coordinates gi3 i = 1,2,3, is a
priori included, only the remainder of the metric tensor

(A.29)

is considered in the following when discussing surface-oriented


coordinates with g-metric (case b of chapter A.2). For surface-
oriented coordinates with a-metric (case c) the surface metric
tensor reads

(A.30)

which is obtained from equation (A.29) by allowing for 9 a ~ !a


on the surface of the body. The components of the g-metric (A.29)
follow from equations (A.4):

(A. 31) k 1,2,3,


a,8 1 ,2 ,

- 208 -
or written explicitly:

(A.32) (131') 2 + (132') 2 + (133') 2


g11 1 1 1

-1' -1 ' -2' -2' -3' -3'


(A.33) g12 g21 B1 B2 + B1 B2 + B1 B2

(A.34) (131' ) 2 + (13;') 2 + (133') 2


g22 2 2

Similarly, with B~'


)
instead of e~', the a-metric is obtained. A
)
notation quite often found in literature on boundary-layer
theory is related to the above a-metric as follows:-

(A.3S)

The off-surface length element ds is given for surface-oriented


coordinates with g-metric by (see equation (A.S»:

(A.36)

and the surface length element with a-metric by:

(ds)2 1 2 1 2 2 2
(A.37) a 11 ( dx) + 2 a 1 2 dx dx + a 22 ( dx )

In appendix A.12 it will be shown how the g-metric can be deter-


mined explicitly from the curvature properties of the surface
which will be discussed next (see also appendix A.14).

- 209 -
A.7 The Covariant Curvature Tensor bas of the Surface

The covariant curvature tensor bas of the surface is defined, for


instance, by [ 71

(A.38) a . a a,S 1,2.


-a,s -3

Like the surface metric tensor the curvature tensor is symmetric,


too

(A.39)
12
baS (""
b 12
b
b 22
)

where because of equations (A.38)

3
(A.40) E
k'=1

A.8 The Two Principal Curvatures of the Surface

The two principal curvaturesK 1 and K2 of the surface are deter-


mined with the help of the components of the curvature and metric
tensor [ 7 , ,1 21 :

a 11 b 22 + a 22 b 11 - 2 a 12 b 12
(A.41) K1 + K2 2
a 11 a 22 - (a 12 )

2
b 11 b 22 - (b 12 )
(A.42) K1 K2 2
a 11 a 22 - (a 12 )

- 210 -
The principal curvature radii are

(A.43) R,
,
K1

(A.44) R2
,
K2

A.9 Directions of the Two Principal Curvatures of the Surface

If Ao = (dx,/dx 2 )o' 0 = ',2, are the directions of ~he principal


curvatures of the surface, they are obtained as the solution of a
quadratic equation [7]:

(A.45)

Another form, which maintains the quadratic equation but which


often is easier to discuss, reads

(A.46)

A.10 The Contravariant Metric Tensor a oe of the Surface

The contravariant metric tensor a oe of the surface is defined


by
(A.47) o = 1,2.

- 211 -
Here 6 u is the Kronecker tensor, that is a uB is the inverse matrix
£
of (A.30). Inversion of (A.30) yields, again, a symmetric matrix

a 11
12
( a )
IA. 48)
a
21 22
a a

21
where a

The contravariant metric tensor gaB off the surface and also the
kl
general contravariant metric tensor g are similarly defined.

A.11 The Mixed-Variant Curvature Tensor b B of the Surface


u

The mixed-variant curvature tensor b B of the surface is obtained


a
by raising one index of baB (note that this is another form of the
Weingarten formula ~3'a b~ ~B):

(A.49) 1 ,2

with the components

11 12
IA.50) b1 b 11 a + b 12 a
1

12 22
(A.51) b2 b 11 a + b 12 a
1

11 12
(A.52) b1 b 12 a + b 22 a
2

12 22
(A.53) b2 b 12 a + b 22 a
2

Note another form of eg. (A.49): bY ( appendix A.14) .


a

The mixed-variant curvature tensor of the surface is not symmetric

- 212 -
(A.54) btl
a

The Gaussian curvature K is the determinant of eq. (A. 54):


K = K, • K2 = det btl, and the mean curvature is given by H (K, +
, 2 a ,
K2 )/ 2 = (b, + b 2 )/ 2. In orthogonal coordinate systems b,
is one of the principal curvatures and b~ the other, if the
directions of the coordinates x' and x 2 fall into the direction
of the principal curvatures.

A.'2 The Metric Tensor gatl and the Curvature Properties of the
Surface

Relations between the metric tensor gatl for surface-oriented


coordinates and curvature properties are obtained after the
definitions of the curvature tensor are introduced in the
expressions which are obtained from the definition of the metric
tensor gatl = ga . gtl with ga given by equations (A.'6):

(A.55)

3 2
+ (x) a
- 3 ,a • ~3,tl

a,a = ',2.
Using instead of the form (A.38) for the curvature tensor baa
another form ( 71:

(A.56) - ~a • ~3,a

- 2'3 -
allows to implement curvature properties into equations (A.55)
(b al' = b l'a ):

3 + a ya b (x 3 )2
(A.57) gas aal' - 2b as x ay b Sa

3 2
aaB - 2 baB x + b ay bYB
(x3)

a, (3, y 1 ,2.

Thus, written out explicitly, equations (A. 57) yield the desired
relations:

3
(A. 58) g 11 a 11 - 2 b 11 x + (b 11 b 1 + b 12 b~) (x 3 )2
1

3
(A.59) g12 g21 a 12 - 2 b 12 x + (b 11 b 1 + b 12 b;) (x 3 ) 2
2

3
(A.60) g22 a 22 - 2 b 22 x + (b 12 b 1 + b 22 b;) (x 3 ) 2
2

In boundary-layer theory it is physically reasonable to stretch


the distance x 3 normal to the surface with the square root of
the reference Reynolds number (see chapter 2). If this stretching
is applied in equations (A.57) while the curvature terms as part
of the geometrical properties of the considered body are made
dimensionless by means of the reference length one obtains (gas'
aas are also dimensionless) :

which demonstrates why in zero or first-order boundary-layer


theory the use of the a-metric is sufficient.

- 214 -
A.13 Christoffel Symbols and Metric Factors

When differentiating vectors expressed in terms of curvilinear


coordinates the corresponding base vectors have to be differentiated
as well as the vector coordinates [ 7,12]:

(A.62)

with

(A.63)

Note that in literature the Christoffel symbols are sometimes


denotes by r~i. The bar on top of the Christoffel symbols indicates
that surface-oriented coordinates with g-metric are meant. In
terms of metric factors and curvature terms the following expressions
result for the Christoffel symbols (ga3 = ga3 0, g33 = g33 = 1):

(A.64) {~}

(A.65) {~}

(A.66) {2 1 2}

(A.67) {~}

(A.68)

(A.69) {/2}

- 215 -
(A.70) a,tl 1 ,2 ,

(A. 71)

(A.72)

(A.73)

(A.74)

(A.75) o o i 1,2,3,

where 9 = det (gij)

The quantities b Q and btl can be considered as curvature properties


a" a 3
of the coordinate surfaces x = const. away from the surface of the
body. However, care must be taken with this interpretation because
of the skewing of the coordinates x 1 , x 2 = const. as discussed in
chapter A.2 • ba" Q and btla as well as k1)
.. can readily be determined
with the help of equations (A. 57) . In the case of surface-oriented
coordinates with a-metric the bars have to be omitted (the g-metric
is taken on the surface at x 3 = 0) and the familiar expressions
[ 71 result:

(A.76)

(A.77)

- 216 -
Further the following relations hold

-,- -2- ,-, 1 -


(A.78) {, ,} + {2 ,} = k" + "2 k22
2g g" -k
k 01,1
01

(A.79) -1 k-
k 01,2
01

(A.80) £2 1 _1_ k
2 2g g'3 it 01,3
01
where

IA.81) k01 = ,rg = [(gll g22 - (g


12
)2)1/2

For the sake of completeness the remaining metric factors are


listed, too:

_ g22
(A.82) k16 9

g'2
(A.83) k17 9

g12
(A.84) k26 9

_ g11
(A.8S) k27 9

(A.86) k31 6 11

(A.87) k32 2£12

(A.88) k33 £22

(A.89) k41 g11

(A.90) k42 2g 12

(A.91) k43 g22

- 217 -
A.14 Shifters

Shifters relate metric quantities, i.e. base vectors, metric tensor


coordinates or CHRISTOFFEL symbols, associated with a general field
point (x 1 , x 2 , x 3 ) to the corresponding quantities at the location
1 2
(x , x , 0) on the surface, and can also be understood to relate
tensor coordinates. Shifters are functions of the properties of the
surface and the wall-normal distance x 3 •

There are two ways to derive the expressions for shifters. One way
is to consider the covariant base vectors [ 3], see eq. (A.16):

(A.92) a-ex - bYex x 3 -Y


a

with the shifter of the first kind [ 3]

(A.93) IS Y - bY x 3
ex ex

ROBERT has shown that an explicit expression can be derived for


the shifter of the second kind which relate contravariant base vec-
tors:

(A.94)

The corresponding expression arises from the condition gCl • gs


ex
c5 S:

(A. 95)

which yields [3

(A.96)

- 218 -
where

(A.97) M det (M a ) =
6

3
1 - bYy x + b (x 3 )2

1 - (K 1 + K2 ) x 3 + Kl K2 (x 3 )2

KA, A = 1,2, are the principal curvatures.

Another way to derive shifters is to consider the surface-parallel


coordinates of a vector, e.g. the velocity vector, which are measured
in terms of general surface-oriented coordinates x a away from the
surface, a = 1,2. It is then possible to transform these coordinates
into those associated with the coordinates x a on the surface x 3 = 0
itself. Thereby an alternative way to compute the shifters is found.
Here the contravariant velocity coordinates are considered. Consider
the surface-parallel component of the velocity, since the wall-nor-
mal component is not changed if the surface-parallel coordinates x a
are changed:

(A.98) v_(a) = va a
oiia
= ~Y a
-y

where the parentheses indicate that the surface-parallel vector


component is regarded and where (A) denotes the velocity coordi-
nates with respect to the surface coordinates. y(a) is first trans-
formed into the reference Cartesian and then into the surface
coordinates:

(A.99)
..
V 1.

and

(A. 100)

Note that ~* Y (a ) 1/2 With eqs. (A.18) and (A.19) one ob-
(yy)
tains from eq. (A.l00):

(A. 101)

- 219 -
The orthogonality relation

(A.l02)

and the Weingarten formula [7,12],

(A.l03) - bY a
a -Y

where the latter is rewritten

(A.l04) 2 y • -3'a
a

are used to replace the terms within the parentheses in eq. (A.l0l):

(A. lOS)

Hence the shifter of the first kind is obtained. From eq. (A.98)
it is evident that both ways are completely equivalent: the kind
of shifter that relates contravariant components coincides with the
kind of shifter that relates covariant base vectors.

- 220 -
Appendix B: Series Expansion of Vector Quantities in
General Coordinates

Let the vector quantity ~ and its derivatives ai ... m F be given

at the point ~O. The three-dimensional Taylor series expansion


about ~O then reads, if ~ denotes a point in the neighbourhood
of ~O:

(B. 1) F (~O) +

N
+ E
n=1

where a.1 denotes the partial derivatives with respect to the


.
direction Xl, and where RT stands for the residual term:

(B. 2) RT
(N+ 1) !

o < 0i < 1,

i 1,2,3.

If contravariant vector coordinates are used the first derivative


of the vector for general coordinates reads [7 1

k k
(B. 3) a.l-F F , i !lk + F !lk,i

and the second derivative

(B.4) alJ-
.. F a.J (a.1-
F)

k k k Fk
F'ij !lk + F, j !lk,i + F'i !lk,j + !lk,ij

- 221 -
Higher derivatives are easily derived in the same manner. The
derivatives of the covariant base vectors ~k are obtained from
the relations [ 71:

(B.5) ~k,i = {knit ~n

where the Christoffel symbols are defined as follows (see appendix


A. 13)

1 nm
(B. 6) -2 g (g lm,
. k + g mk ,l. - gk'l,m ).

As an application consider the series expansion for the velocity


vector about the stagnation point, the results of which were
already given in equations (4.20). At the stagnation point ~s with
va = 0, a = 1,2:

(B.7)

Since the stagnation point solution (chapter 4) yields only first


derivatives, the series expansion will be limited to the leading
terms (N = 1):

(B.8) v' (~) (x


-s
) + •••••

Using the properties of the locally monoclinic surface-oriented


133
coordinate system: 9 = ge3 = 0, g33 = g 33 = 1, one obtains from
equation (B. 3) :

1 k g1l3
(B.9 a) aa ~ .9.1 {v~ a +
2' v (gae ,k + gek,a -gka,e)}
+

1 k 213
+ .9.2 {v~a + 2' v g (gae,k + 9 13k, a - gka, e) } +

+ ~3 {v 3 _ 1 v e gea,3} ,
'a 2

- 222 -
(B.9b)
33 ~ ~1 {V~3 + .12 vll gla
gall,3} +

2 + .12 vll g 2a +
+ ~2 {v'3 gall,3}

3
+ ~3 v'3

Introducing further the symmetry properties of the stagnation-


i a
point solution: v 'Il = 0, i ~ Il: v = 0: a,1l = 1,2: i = 1,2,3,
yields:

1 1a 3 g2a 3
(B. 10) a1 ~ (x )
-s ~1 {v~l + '2 g gla,3 v } + ~2 {~ gla,3 v },

gla 3 1 2a 3
32 ~ (~s) ~1 {~ g2a,3 v } + ~2 {v~2 + '2 g g2a,3 v },

v (x ) 3
33
-s ~3 v, 3

where the right-hand sides of equations (B.l0), including the


covariant base vectors, have to be evaluated at the stagnation
point. Using the boundary-layer stretching procedures, discussed
in chapter 2. and in appendixD, although the length and velocity
component normal to the surface may not always be very much
smaller than the corresponding quantities in the lateral directions,
one obtains in dimensionless form:

(B. 11) v (~)

- 223 -
~3 ~3 ~ -1/2
+ ~3 ~x v'3 Re ref

With equation (A.61), and with the assumption that the metric
properties aal! and bal! are of order unity, one obtains for large
Reynolds numbers ~ ~ a , g 0 ~ a 0 and g 0 3 ~ - 2 b o. There-
a -a a~ a~1/2 a~, a~
fore, neglecting terms of order Re ref and less, equation (B.11)
finally yields equations (4.20) in vector form:

(B.12)

where the familiar abbreviation Aa = va'a (no summation on a) has


been introduced, and where ~1 and ~2 are still base vectors at ~s.

These base vectors can with first-order accuracy be replaced by


those at the station x in question. At least in the case of small
curvature of the coordinate lines this gives sufficient accuracy.

- 224 -
Appendix C: The Vorticity Vector

From the text books on tensor algebra, e.g. [ 7,12], one obtains
the following expression for the vorticity vector ROT~, after
contravariant velocity components have been introduced:

(C. 1) ROT oi
~ 9:i

9:1
(g) ~1 /2 [ (g 3 j v j ) , 2 - (g 2j v j ) , 3] +

+ 9:2
(g)-1/2 [( g V j )'3
1j - (g3jvj)'11 +

+ 9:3
(g)-1/2 [( g V j ) , 1
2j - (g 1j V j ) '2 1 ,

where 9 = det (g .. ), and where the inner brackets with the terms
1J
gkj are readily recognized as covariant velocity components v k •
The vectors 9:i are the usual covariant base vectors for general
coordinates. For locally monoclinic surface-oriented coordinates in
off-surface points equation (C.l) reduces to the following vector
equation (ga3 = 0, g33 = 1, g21 = g12):
(C.2) Rot ~ 9:1
(g)-1/2 [v 3 '2 - 1
(g12 v )'3 - 2
(g22 v ), 3 1 +

(g)-1/2 [(g11 v 1 )'3 2 3


+ 9:2 + (g12 v )'3 - v ,,1 +

+ 9:3
(g)-1/2 [(g'2 v )"
, +
2
(g22 v )"

2
- 1
(g11 v )'2- (g12 v ) '2 1 ,
which in the frame of first-order boundary-layer theory can be
further reduced, reading in terms of the surface metric tensor
aaB (see appendix A) in surface pOints:

i
(C.3) rot ~ ~i w

V 1 2 1
- a'2 '3 - a 22 v'3 +

- 225 -
1 2
- all v '2 - a 12 v '2]

where the symmetry of the metric tensors has been taken into
7
ac ount: a 12= a 21 , b 12 = b 21 . The physical components n*i or
w*1 are obtained by multiplying the contravariant components with
the respective coordinates of the metric tensors (g .. ) 1/2 or
1/2 3 3 3 3 11
(aaa) . Note that n = n* and w = w* for locally monoclinic
coordinates since g33 = 1.

In first-order boundary-layer theory equation (C.3) is valid


throughout the boundary layer because there it is assumed that
the boundary-layer thickness 0 « Ra max with Ra max =0 (1)
and that 0 «IR 'B I with R , = 0 (1). This follows from
a max a max B
a term-by-term inspection of the relations involving 9: a and gaB
(see appendix A) if it is accepted that b aB is a measure of the
local curvature R- 1 of the surface under consideration. The terms
which can be dropped in first-order boundary-layer theory can
readily be detected in'troducing the boundary-layer stretching
into equation (C.3):

(a)-1/2 { -1/2 ",3 1 2


(C.4) rot Y.. ~1 Re ref v'2 + 2 (b 12 v + b 22 v ) -

Re 1 / 2 (a 22 v 2 ,~ + a 12 v 1 , ~) } +
ref

(a)-1/2 1 -1/2 ",3


+ ~2 {- 2 (b 11 v + b 12 v 2 ) - Re ref v , 1 +

- 226 -
(
a 12 ,1 - a 11 ,2 ) V1 (a ) 2} ,
+ + 22 ,1 - a 12 ,2 v

where rot ~ is made dimensionless with v ref /Lref • It is evident


that with the above classical boundary-layer assumptions terms of
order 0 (1) or less can be neglected in comparison with those of
order 0 (Re 1/2).
ref .

(c.s) -1/2
Re ref rot ~

-1/2 1 2
+ ~2 (a) [a 11 v ,~ + a 12 v ,~l

Note, in particular, that the vorticity component in the direction


normal to the wall is small of higher order in first-order bound-
ary-layer theory.

- 227 -
Appendix 0: Dimensions and Boundary-Layer Stretching

To use contravariant variables in fluid mechanics is not common.


Therefore it seems helpful to present the non-dimensionalizing of
dependent and independent variables and the boundary-layer stret-
ching applied to the most important contravariant quantities used.
This is done in form of tables.

The dimensional contravariant quantity q, the dimensionless contra-


variant quantity q. the physical quantity q-* with correct dimen-
sions, the reference quantity qref' and - i f applicable - the metric
quantity cq , with its reference quantity Cq ref , are related to
each other in the following way:

q*
(0.1) -
q
q*
q
q*/qref
c Je c
Cq q
q qref

In the frame of this book it is assumed that the independent


variables are defined such that c q is dimensionless. If this is
not the case, the reference quantity has to take care of that.

In the tables the dimensions are given in terms of the IS-system.


Remember then the coherence J = Pa· m3 when comparing the
dimensions. If the metric properties do not playa role in equa-
tion (0.1) c q is set "one" (this is the case for all thermodynamic
quantities for instance) indicated by "1" in the tables. Vector
quantities are treated according to the different coordinates as
defined in Appendix A, namely general coordinates and locally
monoclinic surface-oriented coordinates. If the latter are needed
on the surface only (e.g. first-order boundary-layer theory) ,
g-metric terms have to be replaced by a-metric terms.

In the table 0.1 flow variables are listed. and in table D.2
geometrical quantities are considered. Concerning geometrical
quantities one has to remember that in this book the Cartesian
reference quantities are introduced as dimensionless lenghts.
The surface parameters xa, of course, in general have no direct
physical lenght properties. Thus the coordinates of the metric

- 228 -
and curvature tensors gij' aaB and b aB etc. and derivatives
thereof are dimensionless as well because of their definitions
[21. In order to gain some insight into the dimensional proper-
ties of some geometrical quantities, it is assumed in tables
D.2 and D.3 that xii-coordinates are dimensional, with lengths
as dimension. Then the coordinates of most base vectors, and
therefore most of the metrical properties are dimensional, too,
as is, remarkably, also x 3 • Equation (A.57), which determines
gaB as a funktion of aaB and of the components of the curvature
tensor

(D.2)

indicates that b-*aB must have the dimension of x 3 , that is that


of a length. Similarly it follows that 6~* must have the di-
mension of (x 3 )-1.

Table D.3, finally, gives an indication how the geometrical


quantities used can be stretched, once the boundary-layer
stretching has been introduced for the surface-normal coordinate
and velocity component (see equations (2.7) and (2.9». This is,
however, not necessary but removes the stretching parameter Re ref
formally from most relations. Stretched quantities are distin-
guished by a tilde. The stretching parameter C s is then introduced
into equation (D.1) as follows:

(D.3) q q

Since c q = 1 for the quantities in table D.3 it is omitted. The


arguments for a possible stretching of b aB (~nd also b~) follow
again from a discussion of equation (D.2). Note that throughout
the present book b aB and b~ are not used as stretched quantities.

- 229 -
In order to compare results obtained from integrating the boun-
dary-layer equations in contravariant form, with those obtained
e.g. in experiment, the predicted quantities have to be transformed
into correctly dimensioned quantities with the help of equations
(2.6a) and (2.7) or by using the tables given in this appendix
-*
solving for q . It is then sometimes of interest to know deriva-
tives with respect to the physical coordinates x*i without having
to perform transformations. This can be done at a general point
(x',x 2 ,x 3 ) by using the definition of a general length element
(A.5) • The general physical length element along the chosen curvi-
linear coordinate line is linked to the element of the correspond-
ing coordinate by means of the corresponding diagonal term of the
metric tensor, resulting in the derivative:

a a (x (i) ) a -1/2
(0.4) --*-.
oX ~ ox*(i) ox i (g(ii)) oxi

where the brackets mean that no summation is to be performed on i.


The operation (0.4) can be performed once again to yield the second
derivative

(0.5)
ax *.~ ax * J.
1 _a_. 1
- "2 g (ii) ax~

Then only the operand with correct dimensions for the operations
(0.4) and (0.5) needs to be expressed in terms of the contra-
variant quantity which can be done with the help of the tables
given.

- 230 -
Table D.1 Flow Variables

-* -
No quantity q q c remarks
qref q
[dimension] [dimension] I
i -
1 .1 velocity catp. v v*i[m/s] vref[m/s] Ig(ii)
-
i = 1,2,3
,
i surface-oriented locally
1.2 velocity catp. va v*a[m/s] vref[m/s] rronoclinic coordinates
I Ira-:- (aa)
(aa) , ~ a = 1,2
I
1.3 v3 v*3 [mls] 1
" " ,
velocity catp. vref[m/s] surface-normal
component
I
IV I i=1,2,3 general coordi-
w i ;;;*i [s -1] - - -1'
1.4 vorticity catp. w v ref/Lref [s ] \ Ig (ii) nates, otherwise
I like 1 .2 and 1 .3
i
I I
1.5 pressure p p[Pa] Pref[pa] 1 Navier-Stokes eqs. I
I
I
I
I I
I - - 2 I
1.6 pressure p p[Pa] Pref(vref ) [Pa] I 1 boundary layer eqs.
I
-p[kg m-3 ]
II - -3
1.7 density P I Pref[kg m ] 1

1.8 tenp:rrature T T[K] Tref [K] 1

specific heat
1.9 at constant c Cp[J/kgK]
-c [J/kgK] 1
pressure p Pref !
Table D.l Flow Variables (ccmtinued)

-* -
No quantity q q c remarks
'Iref q
[dimension] [dinension]

specific heat
1.10 at ccmstant cv cv [J/kgK] -
c [J/kgK] 1
v ref
volume

1.11 effective viscosity in


viscosity !.L ~ [Pa s] ~ref [Pa s] 1
turJ:ulent flows, too
I
heat i effective heat conductivity
1.12 k k [J/smK] 1
conductivi ty kre f[J/smK] in turbo flows, too

IV
W
N 1.13 time t t[s] Lre/Vref [s] 1
--- --- --- -
Table D.2 Geanetrical quantities (assuming xi' ,x 3 to have lengths as dimensions)

No quantity q -
q-* ,c - C remarks
,C q q qref,c qref q
[dimension) [dimension)

cartesian
xi' -*i'
2.1 reference x [m) Lref [m) 1 i = 1,2,3
coordinate

general
2.2 dx i dX* i [m) 1 19 (ii) i = 1 ,2,3
coordinate

surface-oriented ~
(aa) ,
2.3 locally m::moclinic dx a d;{*a [m) 1 a = 1 ,2
coordinate ra-:-:
g (aa)
tv
W
W 3 -*3
II II x x [m) surface-nonnal
2.4 Lref[m) 1
coordinate

ccmpanent of *) -i' =i'


2.5 aj a j [m) L f [m) - i,j = 1,2,3
base vector re

-1' -1' =1' -i' 1=1,2,3


2.6 II II
aa ' a 3 aa [m),a 3 [-) Lref[m) , - -
a = 1,2

metric i,j = 1,2,3


2.7 - - (2) -
tensor gij,aaa gij,aaa m eref [m2 ) a,a=1,2

derivative agij aaae ag ij aaaS i,j,k=1,2,3


2.8 of metric , -- [m2 ) L2 f [m2 ) -
tensor axk , axY axk axY re a,e,y=1,2
---- --- - - - _ .. _ - - - - -_.- -- -- ---_._----
-~--

*) here the overbar denotes ccmpanent of base vector of g-netric


Table 0.2 Gealetrical quantities (assuming xi' ,x 3 to have lengths as dimensions) (continued)

No quantity cq
-
cq -cqref remarks
[dimension] [dimension]

curvature I i,j = 1,2,3


2.9 bij,baB bij ,baB [rn] Lref [rn] I
tensor a,B=1,2

3.0 mixed-variant bB 6 B [rn- 1 ] L


- -1 [rn-1 ]
curvature tensor a a ref a, B = 1,2

3.1 - -1 -1
curvature K Lref[rn ] a = 1,2
N
Ka [rn- 1 ]
a
IN
,c:..

3.2 radius of R
curvature a Ra [rn] Ere fern] a=1,2
--
Table D. 3 Georretrical quanti ties and boundary-layer stretching (cq = 1 for 3. 1 to 3. 5)

~ ~* -
No quantity q q qref Cs remarks
[dimension] [dimension]

3. ",3 -*3 - 1/2 llOnoclinic


3.1 x -coordmate x x [m] Lref[m] (Reref ) coordinates

corrponent of *) -i I -i I -1/2 .
3.2 base vector S'" S3[-] - (Reref ) l.=1,2,3
3

3.3 curvature 15. . 15 j). . j) [m] £ [m] (Re) 1 /2 Ii i , j = 1,2,3


tensor l.J' as l.J' as ref ref S- 1 2
i a, - ,
W
'"
111
3.4 curvature i!la,~ j)a,K [m- 1 ] £-f1 [m- 1 ] (Re f)-1/2
B a Ba re re
II! a,S=1,2
3.5 radius of ~ I R [m] ; £ [m] (Re) 1/2 Ia=1 2
_ _a_t_Ul:'_e_
~_ _ _~_curv a _. a ____ I __~:f _ ref _ l __ ~ ___ . __
*)
see note on Table D. 2 •
Appendix E: Calculation of Streamlines, Surface
and Volume Element

When using contravariant variables the question arises how to


determine the projections of streamlines on surfaces x j = const.,
and, in particular, how to calculate the limiting wall streamlines
on the surface of the body x 3 = o. Streamlines are characterized
by being locally parallel to the velocity vector:

(E.1 ) v x ds o

where ds is a length element of the streamline. For general cur-


vilinear coordinates equation (E.1) results into the expression
[ 7,12]

(E.2) o i,k,l, 1,2,3,

where

1/2

{
g for ikl cyclic 123,

1/2
Eikl = - g for ikl cyclic 132,

0 otherwise.

The evaluation of equation (E.2) yields

(E.3)

or written explicitly

2 3
(E. 4) A1 v dx 3 - v dx 2 0

A2 v
3
dx 1 - v
1
dx 3 0

1 2 1
A3 v dx 2 - v dx 0

- 236 -
By combining equations (E.4) with each other one arrives at the
same well-known form of the equation for the streamline as for
Cartesian coordinates

(E. 5)

Note, that only two of the equations (E.4) which determine the
streamline are linearly independent. It is then obvious that, like
for Cartesian coordinates, the streamlines can be obtained as
common lines of surfaces of two different families of streamfunc-
tions. The projection of streamlines on,.for example, the surface
x 3 = const. requires the integration of the first of equations
(E.5). On the surface of the considered body, x 3 = 0, one has,
of course, to consider as usual the derivatives of v 1 and v 2 with
respect to x 3 since both velocity components vanish (concept of
limiting wall streamlines or skin-friction lines [57]). Note
further that equations (E.5) would result as well if, instead of
covariant components Ai' contravariant components Aj (covariant
base vectors) were introduced in equation (E.2).

Equation (E.5) reads for physical variables

*1 *2 *3
dx dx dx
(E.6)
*1
v
~
v
*3
v

since the ratios dxi/v i remain unchanged when switching from


contravariant quantities to those with correct dimensions (see
appendix D). Consequently, if the dependent variables only are
physical, one obtains

(E.7) (g ) 1/2 ( ) 1/2


22 g33

instead of eq. (E.5).

- 237 -
If, on the other hand, a curve s is given in the surface-oriented
i
x -coordinate system, the directional angles ~ and A, Fig. (E.1)
are found by constructing the true physical relations by means of
3
eg. (A.36), where dx has to be a non-stretched quantity.

If P is a point in the surface oriented xi-coordinate system on


the curve s, then

X 3'
X 2'

X l'
Fig. E.1 Directional angles of a curve s in the
surface-oriented locally monoclinic
i
x -coordinate system .

1 2 3
(E.8) tg ~ (x ,x IX )

Ieg11 (x 1 ,x 2 ,X 3 )dx 1 + cos % (x


123
,x ,X ) •
Icg22(x 1,X2,X3)dx
2

- 238 -
and the elevation angle

123
(E. 9) tg A (x ,x ,x )

A physical surface element dS is found by [7,12]:

r- 1 2
(E. 1 0) dS va dx dx ,

and a volume element dV by :

-
;g 1 2 3
(E. 10) dV dxdxdx.

- 239 -
List of Symbols

The number in parentheses is the page number where the symbol is


defined or where it appears first.

1. Latin Letters

Ai abbreviation, (236)
Aa = v~ a , partial derivative of contravariant velocity
coordinate, (29)
a determinant of covariant surface metric tensor, (15)

a coefficient in scaling relation, (182)

~a covariant base vector of surface-tangential coordinate, (203)

~3 covariant base vector of surface-normal coordinate, (200)

and (204)

ai abbreviation, (168)

a i3 coordinate of covariant surface metric tensor, (14)


-a ref reference speed of sound, (15)
aaa coordinate of covariant surface metric tensor, (208)
aaa coordinate of contravariant surface metric tensor, (211)

B transformation matrix, (205)


b function in scaling relation, (182)
b determinant of mixed-variant surface curvature tensor, (219)

baa coordinate of covariant surface curvature tensor, (210)

b aS coordinate of mixed-variant surface curvature tensor, (212)

C magnitude of largest local surface curvature, (70)


Cr velocity-derivative ratio,(191)
c chord length, (105)

ci constant in boundary-layer thickness relation, (109)

c specific heat at constant pressure, (14) and (231)


p
c general metric quantity, (228)
q

- 240 -
cr velocity ratio, (185)
Cs general boundary-layer stretching parameter, (229)
Cv specific heat at constant volume, (232)

D largest gradient of surface curvature, (70)

D diameter, (133)
E energy quantity, (25)
reference Eckert number, (15)
~i' unit vector of Cartesian reference coordinate system, (199)
~ general vector quantity, (221)
Fk general contravariant vector coordinate, (221)

Fa abbreviation in chapter 5, (41)


f scaling function, (181)
f:f:' ••• = df/dX~, ••• , derivatives of scaling function, (183)

f' v1/v~ , velocity variable, (185)


g determinant of covariant off-surface metric tensor, (216)
and (217)
g = a 12 ' off-diagonal coordinate of covariant surface metric
tensor, (209)
g' 2 2
= v /vr ' velocity variable, (185)

g.
-J
covariant base vector of general coordinates, (199)
gij coordinate of general covariant metric tensor, (200)
gkl coordinate of general contravariant metric tensor, (212)

g3 covariant base vector of off-surface pOints, (204)

~a covariant base vector of off-surface points in locally


monoclinic surface-oriented coordinates, (204)

gv6 coordinate of contravariant off-surface metric tensor,


(212) :
gaB coordinate of covariant off-surface metric tensor in locally
monoclinic coordinates, (208)

- 241 -
H abbreviation in chapter 13, (192)
H mean surface curvature, (213)
h enthalpy per unit mass, (25)

= ~ , (a 22 ' square root of diagonal coordinates of co-


variant surface metric tensor, (209)
K Gaussian curvature, (213)

Kmn coefficients in chapter 13, (186)

Ka principal surface curvatures, (210)

k heat conductivity, (15) and (215)

k mn metric factors, (14) and (232)

L length of body, (131)

reference length, (14)

normalizing length, (119)

M number of increments, (180)


M abbreviation in chapter 10, (136)
M determinant of shifter of the first kind, (75) and (219)

reference Mach number, (15)

shifter of the first kind, (74) and (218)

N abbreviation in chapters 9 and 10, (106) and (126)

o (q) order of magnitude of q, (70)

Po pivotal point in series expansion, (159)

reference Prandtl number, (15)

p static pressure, (14) and (231)

q general velocity, (25)

qi general contravariant vector coordinate, (13)


*3
q physical heat flux, (20)
- -* ,qref
q,q - general dimensional quantities in Appendix D, (228)
'\,

q,q general quantities with boundary-layer stretching in

Appendix D, (229)

- 242 -
R general radius of curvature, (226)
Re ref reference Reynolds number, (15)
ROT rotation in general coordinates, (225J
Rot rotation in off-surface pOints in locally monoclinic
surface-oriented coordinates, (225)
Ra principal surface curvature radius, (211)
R' general position vector, (199)
r radius, (133)
r factor in scaling relation, (182)
rN nose radius of wing cross section or airfoil, (109)
rot rotation in locally monoclinic surface-oriented
coordinates in the frame of boundary-layer theory, (225)

r' surface position vector, (200)


s length along line, (17)
s wing half span, (105)
T static temperature, (14) and (232)
Tij contravariant coordinate of general stress tensor, (80)
t time, ( 1 4 ) and ( 233)
u ratio of velocity derivatives, (191)
u velocity variable, (185)
V ratio of velocity derivatives, (191)
v velocity variable, (185)
y velocity vector, (73) and (219)
vi general contravariant velocity coordinate, (231)
va contravariant surface-tangential velocity coordinate, (231)
v3 contravariant surface-normal velocity coordinate, (231)
v; equivalent inviscid outflow velocity, (22)
w transformed surface-normal velocity coordinate, (186)
xi general coordinate, (199)

- 243 -
xi Cartesian coordinate in chapter 12, (159)
-i
x orthogonal surface-oriented coordinate, (158)
i'
x Cartesian reference coordinate, (199)
xa Gaussian surface parameter, (10) and (200)
x3 surface normal locally monoclinic coordinate, (200)
position vector, (199)

2. Greek Letters

ac contour angle, (120)


i'
flj coordinates of covariant base vector, (199)
j
fli' inverse of fl.i' , (206)
J
n
r ki Christoffel symbol of second kind, (215)
y ratio of specific heats, (15)
abbreviation, (204)
integral quantities, (76)
integral quantities, (77)
integral quantities, (78)
boundary-layer thickness, (17)
61 malls-flow dis,placement thickness, (21)
mass-flow displacement thickness in xa-direction, (22)
momentum-flow displacement thickness in xa-direction, (23)
momentum-flow displacement thickness in x a ,x fI-direction, (24)

energy-flow displacement thickness, (25)


energy-flow displacement thickness in xa-direction, (25)
Kronecker tensor, (212)
error bound, (98)
coordinate of covariant E-tensor, (236)

- 244 -
scaled surface-normal coordinate, (181), (183) and (184)

e vi Ii ' divergence of the velocity vector y, (82)


T
e -- , scaled temperature, (184)
Te
angle between surface coordinate lines, (200)
shifter of the second kind, (218)
second viscosity coefficient, (81)
separation angle, (160) and (161)
streamline-elevation angle, (239)
direction of principal curvature, (211)
viscosity coefficient, (14) and (232)
IJ I bulk viscosity coefficient, (81)
\I =~, cinematic viscosity, (109)
p
locally monoclinic surface-oriented coordinates, (158)
Gaussian surface parameter, (50)
11 ratio of the circumference of the circle to its diameter
p density, (14) and (231)
density of equivalent inviscid outflow, (22)
contravariant wall shear stress coordinate, (19)
dissipation function, (81)
~ function in net-point distribution relation, (181)
~ scaling function, (185)
~LE local leading-edge sweep of general swept wing, (122)
~p local sweep angle of xl = constant-coordinate line, (124)
~o leading edge sweep angle of infinite swept wing, (44)
~ scaling function, (185)
~ angle between x 2 = constant-coordinate line and streamline
of surface-parallel flow, (238)
,..,i i
.. ,1.11 coordinates of vorticity vector, (225)
general scalar function, (207)

- 245 -
3. Indices

3.1 Upper and Lower Indices

i,j,k, •• = 1,2,3; denotes general tensor quantities, parameters,

etc., (199 )

a,B,y, •• = 1,2; denotes surface-parallel tensor quantities,

parameters, etc., (200)

3 denotes surface-normal tensor quantities, parameters,

etc., (200)

3.2 Upper Indices

quantity in Cartesian reference coordinate system, (199)


and (205)
* physical quantity, (228)
+ ++ •• derivate of original expression, (52)

3.3 Lower Indices

A refers to Itl- minimum line, (166)

b body coordinate system, (32)

e external inviscid flow

inv inviscid flow

ip refers to points-of-inflexion line, (169)

LE wing leading edge

1 lower side of wing

m refers to Itl- minimum line, (164)

min minimum

R wing root

ref reference quantity, (228)

- 246 -
S stagnation point
s stagnation-point coordinate system, (32)
T wing tip
TE wing trailing edge
u upper side of wing
w,wall wall or surface

(aa),(ii) diagonal element of a metric tensor, no summation, (230)

(i) , (a) no summation


z refers to p-maximum line, (174)

3.4 Other Symbols

overbar, dimensional quantity, (228), off-surface metric


quantity, (203) and (215)
tilde, quantity with boundary-layer stretching, (229);

quantity in orthogonal wing coordinates, (105)

circumflex accent, shifted quantity, (74) and (219)


underbar, vector quantity, (199)
quantity at point P, (160)
Christoffel symbol of the second kind, (215)
covariant derivative, (75)

'i partial derivative: V't -


i
ar-
av i i
V'j •
ilvi
::J
ilx '
i
v, jk
partial differentiation of vector quantity, (221 )

i=1,2,3:

The Einstein summation convention is uaed throughout, e.g.:

j
Ili' v
i' j
= Ill' v
1' j
+ 1l2' v
2' ... j
1l 3 , v
J'

11 + AQ All a 12 + AQ All a 21 ... Aa All a 22


Aa AS6 aYc') .. A~ All1 a 1 2 2 1 2 2
Y

- 247 -
References

1. Basic Papers

1. KUX, J.: "tiber dreidimensionale Grenzschichten an gekrUrrun-


ten W~nden", doctoral thesis 1971, Univ. of Hamburg, Insti-
tut fur Schiffbau der Universit~t Hamburg, Bericht Nr. 273
(1971).
2. ROBERT, K., GRUNDMANN, R.: "Basic equations for non-reac-
ting Newtonian fluids in curvilinear, non-orthogonal, and
accelerated coordinate systems", DLR-FB 76-47 (1976).
3. ROBERT, K.: "Higher-order boundary layer equat ions for
three-dimensional, compressible flow", DLR-FB 77-36 (1977),
pp. 205-215, also ESA-TT-518 (1979), pp. 273-288.

2. Books

4. ABBOTT, I. H., v. DOENHOFF, A. E.: "Theory of wing


sections", Dover Publications, INC., New York 1959.
5. ARIS, R.: "Vectors, tensors, and the basic equations of
fluid mechanics", Prentice-Hall, Inc., Enqlewood Cliffs
1962.
6. van DYKE, M.: "Perturbation methods in fluid mechanics",
annotated edition, The Parabolic Press, Stanford 1975.
7. KLINGBEIL, E.: "Tensorrechnung fUr Ingenieure",
Bibliographisches Institut, Mannheim, Wien, ZUrich,
Hochschultaschenbuch Band 197, 1966.
8. NASH, J. F., PATEL, V. C.: "Three-dimensional turbulent
boundary layers", SSC Technical Books, Atlanta 1972.
9. RICHTMYER, R. D., MORTON, K. W.: "Difference methods for
initial value problems", 2nd edition, Interscience
Publishers, Inc., New York 1967.
10. ROSENHEAD, L. (ed.): "Laminar boundary layers", Oxford at
the Clarendon Press 1963.
11. SCHLICHTING, H.: "Boundary layer theory", 7th edition,
McGraw-Hill Book Company, New York 1979.
12. SOKOLNIKOFF, I. S.: "Tensor analysis", 2nd edition, John
Wiley and Sons, New York 1964.

- 248 -
3. Boundary Layers

13. ADAMS, J. C. Jr.: "Numerical calculation of the subsonic


and transonic turbulent boundary layer on an infinite yawed
airfoil", AEDC-TR-73-112 (1973).
14. BLOTTNER, F. G.: "Computational techniques for boundary
layers", AGARD-LS-73 (1975), pp. 3-1 to 3-51.
15. BLOTTNER, F. G., ELLIS, M. A.: "Finite-difference solution
of the incompressible three-dimensional boundary-layer
equations for a blunt body", Computers & Fluids, 1 (1973),
pp. 133-158. -
16. CEBECI, T.: "Attachment-line flow on an infinite swept
wing", AlAA J., ~ (1974), pp. 242-245.
17. CESECI, T.: "Progress in the calculation of three-dimensio-
nal boundary layers on bodies of revolution at incidence",
Proc. DEA-Meeting on "Boundary-Layer Effects·, Monterey
1978, AFFDL-TR-78-111(1978), pp. 237-246.
18. CEBECI, T., KAUPS, K., RAMSEY, J. A.: "A general method for
calculating three-dimensional compressible laminar and tur-
bulent boundary layers on arbitrary wings·, McDonnel
Douglas Report J 7267 (1976) (see also NASA CR 2777
(1977».
19. CEBECI, T., KHATTAB, A. A., Stewartson, K.: "Prediction of
the three-dimensional laminar and turbulent boundary layers
on bodies of revolution at high angles of attack".
Proceedings: 2nd Symp. on Turbulent Shear Flows, Vol. II,
pp. 15.8-15.13, Imperial College, London (1979).
20. CESECI, T., KHATTAS, A. A., STEWARTSON, K.: "Studies on
three-dimensional boundary layers on bodies of revolution
II. Three-dimensional laminar boundary layers and the OK of
accessibility", McDonnel Douglas Report J 8716 (1980).
21. COUSTEIX, J.: "Solutions de similitude et methode integrale
de calcul des couches limites laminaires tridimensionelles
sur paroi adiabatique", CERT-Oocument No. 1/5004 ON, Note
Technique (1973).
22. COUSTEIX, J.: "Analyse theorique et moyens de prevision de
la couche limite turbulente tridimensionelle", ONERA Publ.
No. 157 (1974), English translation ESA TT-238 (1976).
23. COUSTEIX, J.: personal communication.
24. EAST, L. F.: "Computation of three-dimensional turbulent
boundary layers", FFA TN AE-1211 (1975).
25. EICHELBRENNER, E. A.: ·Three-dimensional boundary layers",
Annual Review of Fluid Mechanics, Vol. 5 (1973), pp.
339-360.

- 249 -
26. FANNEr.0p, T. K.: "A method of solving the three-dimensional
laminar boundary-layer equations with application to a
lifting reentry body", AIM J., ~ (1968), pp. 1075-1084.

27. FANNEL{IlP, T. K., KROGSTAD, P. A.: " Three-dimensional


turbulent boundary layers in external flows: a report on
Euromech 60", ,1. Fluid Mech., l! (1975) pp. 815-826.

28. GEISSLER, W.: "Calculation of the three-dimensional laminar


boundary layer around bodies of revolution at incidence
and with separation", AGARD CP-168 (1975), pp. 33-1 to
33-11.

29. GRUNDMANN, R.: " Grenzschichttheorie zweiter Ordnung fur


stationare, laminare, kompressible Stromungen an gekrummten
Oberflachen", doctoral thesis, TU Berlin, 1972, also DLR-FB
73-31 (1973), English translation ESRO TT-85 (1974).

30. GRUNDMANN, R: "A non-orthogonal coordinate system for


calculating boundary layers along lines of symmetry", Proc.
DEA-Meeting on "Viscous and Interacting Flow-field
Effects", Annapolis 1980, AFFDL-TR-80-30R8 (1980), pp.
347-358.

31. GRUNDMANN, R., ROBERT, K.: "Krummungseinflusse auf das


Abloseverhalten laminarer Grenzschichten", DLR-FB 77-36
(1977), pp. 169-177, English translation ESA-TT-518 (1979).

32. HANSEN, A. G.: "Compressible, three-dimensional, laminar


boundary layers - a survey of current methods of analysis",
Douglas Paper 3105 (1964).

33. HEAD, M. R., PRAHLAD, T. S.: "The boundary layer on a plane


of symmetry". The Aeronautical Quartely, Vol. XXV, Pt. 4
(1974), pp. 293-304.

34. HIRSCHEL, E. H.: "The influence of the free-stream Reynolds


number on transition in the boundary layer on an infinite
swept wing", AGARD-Report No. 602 (1973), pp. 1-1 to 1-11.

35. HIRSCHEL, E. H.: "Boundary-layer equations in holonomic


formulation", Proc. Int. Conf. on Numerical Methods in
Laminar and Turbulent Flow, Pentech Press, London 1978, pp.
421-432.

36. HIRSCHEL, E. H., JAWTUSCH, V., GRUNDMANN, R.: "Berechnung


dreidimensionaler Grenzschichten an Pfeilflugeln", DGLR
paper 76-187 (1976), DGLR-Jahrbuch, Koln 1977.

37. HIRSCHEL, E. H., ~1AWTUSCH V.: "Nachrechnung des experimen-


tell ermittelten Ubergangs laminar-turbulent an einem ge-
pfeilten Fluqel", DLR-FB 77-16 (1977), pp. 179-190.

38. IHRSCHEL, E. H., SCHWAMBORN, D.: "Ein Verfahren zur Berech-


nung von Grenzschichten in Stromunqssymmetrieebenen",
DLR-FB 77-16 (1977), pp. 125-132.

- 250 -
39. HIRSH, R. S., CEBECI, T.: "Calculation of three-dimensional
boun~ary layers with negative cross flow on bodies of revo-
lution", AIAA paper 77-683 (1977).

40. HmvARTH, F. R. S.: "The boundary layer in three-dimensional


flow-Part II. The flow near a staanation point",
Philosophical Maqazine, Ser. 7, vol. 42, No. 335 (1951),
pp. 1 433 -1440 .

41. KITCHENS, C. W., JR., SEDNEY, R., GERBER, N.: "The role of
the zone of dependence concept in three-dimensional
boundary-layer calculations", USA Ballistic Research
Laboratories, Aberdeen Proving Ground, Maryland, R 1821
(1975).

42. KORDULLA, 1'1.: "Inviscid viscous interaction in transonic


flows about finite three-dimensional winge", AIAA J., 1~
(1978), pp. 369-376 (AIAA paper 77-209 (1977)).

43. KORDULLA, W.: "Viscous transonic flows about 3-D winas",


Z FW, 3. (1 9 7 8), pp • 3 0 6 - 3 1 2 •

44. KRAUSE, E.: "Comment on "Solution of a three-dimensional


boundary-layer flow with seperation"", AIAA cTollrnal, 7
(1969), pp. 775-776.

45. KRAUSE, E.: "Numerical treatment of boun~ary-Iayer pro-


blems", AGARD LS-64, Brussels (1973), Pp. 4-1 to 4-21.

46. KRAUSE, E.: "Analysis of viscous flow over swept winqs",


ICAS Paper No 74-20 (1974).

47. KRAUSE, E., HIRSCHEL, E. H., BOTHMANN, TH.: "Numerische


Stabilit~t dreidimensionaler Grenzschichtstr6munaen", ZAMM
Sonderheft GAr1M-Taaung, Band 48 (1968), pp. T205-T20S.

48. KRAUSE, E., HIRSCHEL, E. H., BOTHMANN, TH.: "Die Numerische


Integration der Bewegungsgleichungen dreidimensional~r,
laminarer, kompressibler Grenzschichten", DGLR-Fachbuch-
reihe, Band 3, Braunschweig 1969, pp. 03-1 to 03-49.

49. KRAUSE, E., HIRSCHEL, E. H., BOTHMANN, TH.: "Differenzen-


formeln zur Berechnunq dreidimensionaler Grenzschichten",
DLR-FB 69-66 (1969).

50. KRAUSE, E., HIRSCHEL, E. H., KORDULLA, 1'1.: "Fourth-order


"Mehrstellen"-inteqration for three-dimensional turbulent
boundary layers", Proc. AIAA Compo Fluid Dynamics Conferen-
ce 1973 (see also Computers and Fluids, 4 (1976), pp.
77-92).

51. KREPLIN, H. P., VOLLME~S, H., MEIER, H. D.: "Experi~ental


determination of wall shear stress vectors on an inclined
prolate spheroid", Proc. DEA-Meetina on "Viscous and
Interacting Flow-field Effects", Annapolis 1980,
AFFDL-TR-80-3088(1980), pp. 315-332.

- 251 -
52. KROGSTAD, P. A. : "Investigation of a three-dimensional
turbulent bounoary layer driven by simple two-dimensional
potential flow", doct. thesis, Norwegian Institute of
Technology, Trondheim 1979.

53. LIBBY, P. A.: "Heat and mass transfer at a general three-


dimensional stagnation point", AIM ,lollrnal, 5 (1967), pp.
507-517.

54. LIBBY, P. A.: "Three-oimensional boundary layer with uni-


form mass transfer", The Physics of Fluids, 12 (1969) pp.
408-417.

55. LIBBY, P. A., CHEN, K. K.: "Boundary layer at a swept


staqnation line of semi-infinite span", J. Fluid Mech., 41
( 1 970), pp • 7 37- 750 .

56. LIGHTHILL, M. J.: "On displacement thickness" J. Fluid


Mech., ! (195B), pp. 383-392.

57. LIGHTHILL, M. J.: "Introduction boundary layer theory", in


"Laminar boundary layers", L. Rosenhead (ed.), Oxford at
the Clarendon Press 1963, pp. 46-113.

58. LINDHOUT, J. P. F., MOEK, G., de BOER, E., van den BERG,
B.: "A method for the calculation of 3-D boundary layers on
!)ractical wing configurations", The Joint ASME-CSME Applied
Mechanics, Fluid Engineering ano Bioengineering Conference,
Niagara Falls, New York, June 18-20, 1979 (NLR MP 79003 U).

59. LINDHOUT, J. P. F., van den BERG, B., ELSENAAR, A.: "Com-
parison of boundary-layer calculations for the root section
of a wing: the September 1979 Amsterdam Workshop test
case", NLR-report to be published.

60. MAGER, A.: "Generalization of boundary-layer momentum-inte-


gral equations to three-dimensional flows including those
of rotating systems", NACA Report 1067 (1952).

61. McLEAN, J. D.: "Three-dimensional boundary-layer calcula-


tions for swept wings", AIAA Paper 77-03 (1977).

62. McLE:AN, J. D., RANDALL J. I,.: "Computer program to calcu-


late three-dimensional boundary-layer flows over wings with
wall-mass transfer", NACA CR 3123 (1979).

63. MOORE, F. K.: "Displacement effect of a three-dimensional


boundary layer", NACA Rep. 1124 (1953.

64. MORKOVIN, M. \T.: "Instability, transition to turbulence,


and predictability", AGARD-AG-236 (1978).

65. MORKOVIN, M. V.: "Technical evaluation report of the Fluid


Dynamics Panel Symposium on Laminar-Turbulent Transition,
Copenhagen, 1977", AGARD-AR-123 (1978).

- 252 -
66. NASH, J. F., SCRUGGS, R. M.: "An implicit method for the
calculation of three-dimensional boundary layers on finite,
thiCk winqs", SYBUCON Report, SYB-76-102, Atlanta (1976).

67. NASH, J. F., SCRUGGS, R. M.: "Verification of a three-


dimensional turbulent boundary-layer calculati.on method",
AFFDL-TR-7~-15 (1978).

68. PAPP8NFUSS, H. D.: "Die Grenzschichteffekte zweiter Ordnunq


am dreidimensionalen Staupunkt mit starkem Absauqen oder
Aushlasen", ZFW,.!. (1977). op. 87-96.

69. PH~H.S,w. J., SCHIl:'HOL'f, G ••J., van r1en BERG, 13.: "Calcula-
tion of the flow around a swept wing taking into account
the effect of the three-dimensional boundary layer", NLR TR
75076 (1975).

70. Poll, o. 1. A.: "Some aspects of the flow near a swept


attachment line with particular reference to boundary-layer
transition", doctoral thesis, Cranfield, C. of A. Rep.
7805./L (197R).

71. PRANOTL, L.: ROber Reihungsschichten bei dreidimensionalen


Stromunqen", Festschrift zum 60. Geburtstage von A. Betz,
Gottingen 1945, pp. 134-141.

72. RAETZ, G. S.: "A method of calculating three-dimensional


laminar boundary layers of steady compressible flows",
Northrop-Aircraft Inc., Rep. No. NAI-58-73 (BLC-114),
( 1957) .

73. RUBESIN, M. W.: "Developments in the computation of tur-


bulent boundary layers", NASA TM-78620, (1979).

74. SOIMIT'r, V., COUSTEIX, J.: "Etude de la couche limite tri-


dimensionelle sur une aile en fleche", ONF.:RA Rapport
Technique No. 14/1713 AN (1975).

75. SCHNRIDER, G. R.: "Die Berechnunq r1er turbulenten Grenz-


schicht am schiebenden Flugel unendlicher Spannweite mit-
tels eines dreidimensionalen Mischungswegmodells", DLR-FB
77-73 (1977). .

76. SCHi~EIDER, G. R.: "Calculation of three-ciimensional boun-


dary layers on bodies of revolution at incidence", Proc.
DEA-Meeting on "viscous and Interacting Flow-field
Effects", Annapolis 1980, AFFDL-TR-80-3088 (1980), pp.
287-314.

77. SCIIONAUER, W., GRUNING, J. GLOTZ, G., OAUBLER, H. G.:


"Selfadaptive solution of the incompressible 3-D boundary-
layer equations with controlled error", Proc. 2. GAMM-Conf.
on Num. Meth. in Fluid Mechanics (E. H. Hirschel,
W. Geller, eds.), Koln 1977, pp. 184-191.

- 253 -
78. SCHWAMBORN, D.: "Boundary layers near the attachment line on
finite wings and related bodies at angle of attack", 15th
IC~AM, ~oronto, Canada, Aug. 1980 (to appear as doctoral
thesis, RWTH Aachen).

79. SMITH, P. D.: "An integral prediction method Eor


three-dimensional compressible turbulent boundary layers",
ARC R. & M. No. 3739 (1974).

80. SQUIRE, L. C.: "~he three-dimensional boundary-layer


equations and some power series solutions", ARC Techn. Reo.
R. & M. No. 3005 (1957).

81. STOCK, H. W.: "Integral method for the calculation of


three-dimensional laminar and turbulent houndary layers",
NASA-TM-75320 (1978).

82. STOCK, H. W.: "Laminar bondary layers on inclined


ellipsoids of revolution", ZFW, ! (1980) pp. 217-224.

81. TASSA, Y., RESHOTKO, E., ANDERSON, B. H.: "Finite


riifference proceriure for boundary layers incillriing effects
of longitudinal and transverse cllrvatures", AIAA-paper
76-427 (1976).

84. TREADGOLD, D. A., BEASLEY, J. A.: "Some examples of the


application of methods for the prediction of boundary-layer
transition on sheareil wings", AGARD-Report No. 602 (1973),
pp. 2-1 to 2-11.

1:l5. 'ERELLA, M., LIBBY, P. A.: "Similar solutions for the hyper-
sonic laminar boundary layer near a plane of symmetry",
AIM .1., 1 (1965), pp. 75-81.

86. WANG, K. C.: "Three-dime'1sional laminar boundary layer 'wer


a body of revolution at incidence, Pt. II. The role of
subcharacteristics and improved method of calculation",
Martin Marietta Corp. Research Institute for Advanced
Studies, TR 69-13 (1969).

87. WANG, K. C.: "Three-dimensional boundary layer near the


plane of symmetry of a spheroid at incidence", J. Fluid
Mech.,.!1 (1970), pp. 187-209.

88. WANG, K. C.: "On the determination of the zones of


influence and dependence for three-dimensional boundary-
layer equations", J. Fluid Mech., i~ (1971), pp. 397-404.

89. WESSELING, P.: "The calculation of incompressible three-


dimensional boundary layers", Part 1: "Formulation of a
system of equations", NLR Rep. No. A~-06-01 (1969).

- 254 -
90. BRO\~, S. N., STEWARTSON, K.: "Laminar separation", in: An-
nual Review of Fluid Mechanics, Vol. 1 (1969). pp. 45-72.

91. [IAN, 1'., PATEL, V. C.: "Flow separation on a spheroid at


incidence", J. Fluid ~lech., 92 (1979), pp. 643-657.

92. HIRSCHEL, E. H., KORDULLA, W.: "Local prooerties of three-


dimensional separation lines", ZFW, ! (1980), pp. 295-307.

93. KRAEMER, K.: "On vortex-sheet separation, with particular


reference to the delta-wing leading-edge separation", paper
presented at the IUTAM Symposium on Concentrated Vortex
Motions in Fluids, Ann Arbor, 1964, also AVA 64 A 17
( 1964) .

94. LEGENDRE, R.: "S~paration de l'ecoulement laminaire tri-


dimensionnel", La Recherche A~ronautique No. 54, Novembre-
D~cembre 1956, pp. 3-8.

95. L~~GENDRE, R.: "Lignes de courant d' un ecoulement permanent


decollement et separation", La Recherche Aerospatiale, No.
6 (1 97 1 ), pp. 327 - 335 .

96. LEGENDRE, R.: "D~collement d'un ecoulement Ie long d'une


ligne d'une surface requliere", La Recherche A~rospaliale,
No. 5 (1978), pp. 21 3 - 219 •

97. LEGENDRE, R.: "Types de decollements de l'ecoulement trans-


sonique autour d'une aile en fleche", L'A~ronautique et
L'Astronautique, Vol. 75, No.2 (1979), pp. 39-42.

98. t1ASKELL, E. C.: "Flow separation in three dimensions", RAE


Rep. No. AERO 2565 (1955).

99. Meier, H. u., KREPLIN, H. P.: "Experimentelle Untersu-


chunqen von Ablosephanomenen an einem rotationssymmetri-
schen Korper", DGLR Symposium: Stromungen mit Ablosung,
19./20. Sept. 1979, Munchen, DGLR paper 79-071 (1979).

100. OSWATITSCH, K.: "Die Ablosebedingungen von Grenzschichten",


in: Boundary Layer Research, H. GORTLER (ed.), Sprinqer-
Verlag, Berlin/Gottingen/Heidelberg 1958, pp. 357-367, also
as NASA TTF-15200 "The separation conditions of boundary
layers".

101. PEAKE, D. J.: "Controlled and uncontrolled separation in


three dimensions", NRC Canada No. 15471 (1976).

102. PEAKE, D. J., TOBAK, M.: "Three-dimensional interaction and


vortical flows with emphasis on high speeds", AGARDograph
No. 252 (1980).

103. vIANG, K. C.: "Separating patterns of boundary layer over an


inclined body of revolution", AIM J., 10 (1972), pp.
1044-1050.
- 255 -
104. MASTIN, C. W., THO~PRON, J. P.: "Transformation of three-
dimensional regions onto rectanqular regions by elliptic
systems", Numerische Mathematik 29 (1978), pp. 397-407.

105. THOMPSON, J. F., THAMES, F. C., MAS'rIN, C. W.: "30unc'lary-


fitted curvilinear cOQrdinate system for solution of par-
tial differential equations of fielc'ls containinq any number
of arbitrary two-dimensional bodies", NASA CR-2729 (1977).

106. YU, N. J.: "Grid qeneration and transonic flow calculation


for three-dimensional confiqur<ltions", THAll. paper 80-1391
(1980) •

6. General Viscous Flow

107. AGERWAL, R., RAKICH, ,T. V.: "Computation of hypersonic,


laminar viscous flow past spinning sharp and blunt cones at
high angle of attack", AIAA paper 78-65 (1978).

108. BALDWIN, B. S., LOMAX, fl.: "Thin layer approximation and


algebraic model for separated turbulent flows", AIAA paper
78-257 (1978).

109. BRILEY, W. R., McDONALD, H.: "Analysis and computation of


viscous subsonic primary and secondary flows", AIM paper
79-1453 (1979).

110. BUGG~LN, R. C., McDONALD, H., KRESKOVSKY, J. P., LEVY, R.:


"Computation of three-dimensional viscous supersonic flow
in inlets", AIAA-paper 80-194 (1980).

111. CHAPMAN, D. R.: "Computational aerodynamics - development


and outlook", AIAA .1.,.!l. (1979), pp. 1293-1313.

112. DEIWERT, G. S.: "Computation of separated transonic


turbulent flows", AIAA .T., ]..! (1976), pp. 735-740.

113. HELLIWELL, W. S., DICKINSON, R. P., LUBARD, S. C.: "Viscous


flow over arbitrary geometries at high angle of attack",
AIAA Paper 80-0064 (1980).

114. HUNG, C. M., MacCORMACK, R. W.: "Numerical solution of


supersonic laminar flow over a three-dimensional
compression corner", AIM Paper 77-694 (1977).

115. HUNG, C. M., MacCORMACK, R. W.: "Numerical solution of


three-dimensional shock-wave and turbulent boundary-layer
interaction", AIAA paper 73-161 (1978), also AIAA J .• 16
(1978), pp. 1090-1096.

- 256 -
116. HUNG, C. M.: "Numerical solution of supersonic laminar flow
over an inclined body of revolution", AIM paper 79-1547
(1979), also AIAA J., ~ (1980), pp. 921-928.

117. KU'I'LER, p., CHAKRAWARTHY, S. R., LOMBARD, C. P.:


"Supersonic flow over ablated nosetips using an unsteady
implicit numerical procedure", AIA.\ paper 78-213 (1978).

118. KUTLER, p., PEOELTY, J. A., PULLIAM, T. q.: "Supersonic


flow over three-dimensional ablated nosetips usinq an
unsteady implicit numerical procedure", AIM paper 80-0063
( 1980) •

119. LEVY, R., McDONALD, H., BRILEY, W. R., KRESKOVSKY, J. P.:


"A three-dimensional turbulent compressil)le subsonic nuct
flow analysis for use with constructed coordinate systems",
AIA1'. paper 80-1398 (1980).

120. LIN, T. C., RUBIN, S. G.: "Viscous flow over a cone at


moderate incidence. I. Hypersonic tip region", Computers
and Fluids, 1 (1973), pp. 37-57, "II. Supersonic boundary
layer", ,1. Fluid Mech., 5q (1973), PP. 593-620.

121. LUBARD, S. C., HELLIWELL, W. S.: "Calculation of the flow


on a cone at high angle of attack", AIM .1.,17 (1974), ~p.
965-974.

122. McRAE, O. S., HUSSAINI, M. Y.: "Numerical simulation of


supersonic cone flow at high angle of attack", AGARD FOP
Symposion on Hiqh Angle-of-attack Aerodynamics, AGARD
CP-247 (1978), pp. 23-1 to 23-10.

123. NIETUBICZ, C. J., PULLIAM, T. H., STEGER, J. L.: "Numerical


solution of the azimuthal-invariant thin-layer Navier-
Stokes equations", AIAA paper 79-0010 (1979).

124. PATANKAR, S. V., SPALDING, D. B.: "A calculation procedure


for heat, mass, ann momentum transfer in three-dimensional
parabolic flows", International Journal of Heat ann Mass
transfer, 22 (1972), pp. 1787-1Q06.

125. PULLIAM, T. H., STEGP.R, J. T.... : "On implicit finite-nif-


ference simulations of three-dimensional flow", AIM paner
7f1-10 (1978), also AIM J., ~ (1980), pp. 159-167.

126. PULLIAM, T. H., LOMAX, H.: "Simulation of three-dimensior'lal


compressible viscous flow on the ILLIAC IV computer", ;\IAA
paper 79-206 (1979).

127. ROBER'rS, D. W., FORESTP.R, C. K.: "A parabolic computational


procedure for three-dimensional flows in ducts with arbi-
trary cross-sections", AIM paper 78-143 (1978), also AIAA
J., 12 (1979), pp. 33-40.

128. SCHIFF, L. B., STEGER, J. L.: "Numerical simulation of


steady supersonic viscous flow", AIM paper 79-13'J (1979).

- 257 -
129. SHANG,.T. 5., HANKEY, ,JR., W. L.: "Numerical solution of
the Navier-Stokes equations for a three-~imensional cor-
ner", AlAI'. J.,..!..2 (1977), pp. 1575-1582.

130. STEGER, J. L., BAILEY, H. E.: "Calculation of transonic


aileron buzz", AIM paper 79-134 (1979).

131. VIGNERON, Y. C., RAKICH, J. V., TANNEHILL, J. C.:


"Calculation of supersonic viscous flow over delta winqs
with sharp subsonic leadinq edqes", AIM paper 78-1137
( 1978) •

132. VIVIANO, H.: "Formes conservatives des equations de la


dynamique des gaz", La Recherche Aerospatiale No.1 (1974),
pp. 65-68.

133. WALITT, L., TRULIO, ,J. G.: "I'. numerical metho~ for compu-
ting three-dimensional viscous supersonic flow fields about
slender bodies", NASA CR-1963 (1971).

- 258 -
Subject Index

Aircraft configuration, 62 ff, 117


Angle of attack, 34 ff, 56, 65 ff, 103, 111 ff, 120, 130, 136 ff,
156, 164, 172, 175, 184
Attached flow, 1 ff, 62, 156
Attachment line, 7 ff, 34 ff, 48, 56 ff, 109 ff, 120, 139, 156 f,
172 ff, 189
curvature of, 173 ff
Base vector, 11 ff, 33, 73, 121 ff, 131, 135 f, 142, 154, 199 ff,
203 ff, 218 f, 222 ff, 225 f, 233 ff, 236 f
Blasius transformation, 180 f
Blowing, see Boundary-layer blowing
Blunt body in supersonic flow, 28
Body, also Fuselage, 1, 4 ff, 10, 23, 32 f, 34 ff, 62, 66,
101 ff, 117, 129 ff, 157 f, 172, 184
axisymmetric (body of revolution), 90, 133 ff, 139 ff
cross section, 129 ff
nose, 66, 129, 138 ff
surface, 4, 10, 17 f, 20, 26, 32 ff, 62, 65 f, 117 ff, 158, 161
Boundary conditions, 4, 17 f, 23, 38, 43, 45, 49, 53, 58, 89 ff,
190
Boundary layer
blowing, 3, 17, 22, 26
cooling, 3, 17 f, 28
heating, 3, 17 f, 26, 28
outer edge, 17, 29, 38, 45, 47, 58, 97 ff, 180, 189
prediction, 1 ff, 14, 34, 62 ff, 180
profile, 22 ff, 29, 63, 150, 180
scaling, 14, 180 ff
stretching, 13 f, 19, 92, 98, 118, 151, 185, 214, 223, 228 f
suction, 3, 17, 22, 26
thickness, 4, 12, 70, 78, 97, 101, 109 ff, 179, 180 ff, 202,
226
Boundary-layer equations, 1 f, 11 ff, 22, 25, 26 ff, 34 ff, 62 ff,
70 ff, 91, 117, 138 ff, 180 ff, 187, 190 ff, 229, 231
first order, 8, 10 ff, 26 ff, 34 ff, 70, 98, 100 ff, 202 f
higher order, 7 ff, 11, 70 ff, 91, 100 ff
Boundary-layer integral parameter, 7 f, 21 ff, 26, 32, 73 ff, 96 ff
(massflow-) displacement thickness, 8, 14, 21 ff, 30, 40, 46, 50,
54, 59, 75 f, 97 f, 184, 189 f, 192 ff
energy-flow displacement thickness, 8, 21, 25, 31, 41, 47, 51, 56,
60, 75, 77, 190
energy-loss thickness, 8, 21
momentum-flow displacement thickness, 8, 21, 23 ff, 31, 40 f, 46,
51, 55, 59 f, 75 f, 190
momentum-loss thickness, 8, 21
Boundary-layer theory, 62 f, 65, 155, 201 f,.209, 214
first order, 10 ff, 19 ff, 24, 44, 48, 70, 78, 100 ff, 110 f,
113 ff, 177, 202 f, 226 f
higher order, 9, 70 ff, 100
second order, 26, 71, 100 ff, 112, 114 ff
zero order, 19 ff, 24, 100 f, 113 ff, 214
Boundary values, see Boundary conditions
Bulk viscosity, 81, 93, 149

- 259 -
Characteristics, 63 f
Characteristic properties, 8, 25, 63 f, 68, 89
Chord, see wing
Chord direction, 67, 69
Christoffel symbol, 80, 86, 92, 215 ff, 218, 222
Circumferential direction, 66
Compatibility conditions, 9, 146 ff, 159
Computational domain or area, 4, 18, 69, 120, 139, 180 ff, 202
Computation method, 2, 180 ff
Continuity equation, 11, 15, 28 f, 37, 45, 49, 52, 58, 71, 80, 82,
87, 92, 146 f, 151, 185 f, 191, 193, 195 ff
Contour angle, 7, 9, 44, 48, 107, 113, 117, 120 ff
Convective terms, 29, 63, 147
Coordinate line, 4, 14, 200 ff, 230
Coordinate systems, 4, 11, 80, 87 f, 117, 199 ff, 221 ff, 225, 228,
231, 236
accelerated, 11
Cartesian, 13,44,87,94,159,173,237
Cartesian reference, 4 ff, 10, 33, 103, 107, 117, 120, 130, 133,
153, 199, 203, 219, 228, 233
external streamline, 7, 65 f, 129 f, 184
orthogonal, 22, 29, 158, 213
percent line, 7, 119
surface oriented, locally monoclinic, 4 ff, 10 ff, 17, 26 ff,
32 f, 35 ff, 44 ff, 66, 71 ff, 82, 87 ff, 103, 107,110,
117 ff, 146, 158 f, 199, 200 ff, 222, 225 f, 228, 231 ff, 238
Corner flow, 6, 94 f, 117
Courant-Friedrichs-Lewy condition, 63, 65, 69, 169
Crocco relation, 182
Cross flow, 65, 69
Cross-flow profile, 176, 180
Curvature of surface, 12, 44, 70, 73, 79, 100 ff, 109 ff, 125 ff,
209 ff, 224, 226, 234 f
direction of principal curvature, 211
Gaussian, 213
gradient, 70 f
mean, 213
principal, 70, 75, 136 f, 143, 210 f, 213, 219
principal radii, 211
radius, 4, 12, 70, 100, 110, 116, 202, 234 f
Curvature tensor of surface
covariant, 126, 136, 143, 151, 154, 210, 213,216 ff, 228 ff, 233 ff
mixed variant, 127, 137, 143, 212 f, 216 ff, 229, 233 ff
Density, 14, 231
Derivative, 11, 29, 38, 230, 233, 237
covariant, 75, 81, 83, 147
of vector quantity, 215, 221 f
partial of scalar, 207
spanwise, 42 ff, 166, 174
Displacement thickness, see Boundary-layer integral parameter
Dissipation function, 81, 85
Dividing stream surface, 34
Domain of dependence, 8, 63 ff, 169
of influence, 8, 63 ff
Drag, 3
Duct flow, 95, 128

- 260 -
Eckert number, 15
Ellipsoid, 133 ff
Energy
equation, 12, 16, 28 ff, 38, 45, 49, 53, 58, 72, 80 ff, 87, 92,
146, 149, 152, 188, 192 ff, 196 ff
flux, 25, 75
kinetic, 25
Energy-loss thickness, see Boundary-layer integral parameter
Enthalpy, 25
Entropy, 28
Equation of state, 15, 93
Equivalent inviscid source distribution, 8, 23, 31, 40, 46, 50, 54,
59, 78, 190 ff, 198
Euler equation, 1, 5, 91, 94
External inviscid flow of boundary layer, 1, 4, 13, 17, 22, 38,
42 ff, 58, 63, 65, 68,75 ff, 109, 114 ff, 138 f , 157, 172 ff,
180 ff, 196

Finite-difference molecule, 65 ff, 169


box scheme, 186
characteristic box, 65, 139
four-point scheme, 67, 69
Lasoonen, 65
z-shaped, 65 ff, b9
Finite-difference solution, 2 f, 34, 39, 89
conditional stability, 65
implicit, 66
marching technique, also direction, 65 ff, 89, 170
net generation, 200
net-point number, 180 ff
net spacing, 6, 180 ff
rate of convergence, 89
stability, 65
Flat-plate flow, 181
Flow
supersonic, 28, 79, 90
transonic, 98, 170
Flow (boundary layer)
laminar, 3,12,14,20,43,45,80,88 ff, 103 ff, 111 ff, 160 f,
180 ff
quasi-one-dimensional, 8, 26 ff, 34, 56, 197
quasi-two-dimensional, 8, 28, 34 ff, 157, 176 f
three-dimensional, 1 ff, 11 ff, 21 ff, 34, 42, 63, 65, 177,184 ff
transition laminar-turbulent, 2 f, 56, 88 ff, 100, 114 f, 181 f
turbulent, 2 f, 12 ff, 20, 43, 88 f, 100, 112, 114, 159, 180 ff,
232
two-dimensional, 1, 21 f, 46, 50 f, 56, 146, 202
Flow-field prediction, 1 f, 32, 34, 66, 118
Fuselage, see Body

Gasdynamic equations, 11
Gaussian surface parameter, 7, 10, 82, 103, 117 ff, 203 ff, 223
Geometrical singularity, 66, 130, 135

Heat conductivity, 14, 232


apparent turbulent, 20
effective, 20, 232
molecular, 20

- 261 -
Heat flux, heat transfer, 18, 20, 26, 30, 39, 46, 50, 54, 59, 73,
95

Independence principle, 45
Infinite swept wing or cylinder (see also locally infinite swept
wing), 34, 42 ff, 90, 103, 108 ff, 128 f, 162, 166 f, 174 f
Initial conditions, also data, 18, 32, 34, 38, 42 f, 56 ff, 60,
b2 ff, 89, 93, 114, 139, 196 ff
Inlet of propulsion engine, 2
Integral method or solution, 2, 21
Integral parameter, see Boundary-layer integral parameter
Interpolation of data, 6, 139, 202
Inviscid flow, see also External inviscid flow of boundary layer,
1, 21, 23, 42, 52, 65 ff, 96, 98
Irrotational flow, 42, 79, 98, 174
Isobar, 42, 52
Inviscid-viscous coupling, 1, 23, 90
Jacobian transformation matrix, 206
Kronecker tensor, 212
Laminar flow, see Flow (boundary layer)
Leading edge, see Wing
Leading edge contamination, 100, 112, 114
Leading-edge oriented coordinates, 44 ff, 48 ff
Length element of coordinate line, 17, 200, 209, 230
Length, physical, 14, 125, 199, 223, 229
Levy-Lees transformation, 180
Limiting wall-streamline, see Skin-friction line
Locally-infinite-swept-wing approximation, 42 f, 52 ff, 56 ff, 67,
69, 195 ff
Logarithmic wall region, 181
Main-stream direction, 66
Metric factor, 15,23, 36 f, 44, 48, 53, 60, 71, 92 ff, 113, 154,
185 f, 215 ff
Metric properties, 4, 7, 70 f, 91, 96, 117, 120 ff, 133 ff, 199 ff
228 ff, 233 f
Metric tensor of locally monoclinic surface-oriented coordinate
system
contravariant, surface, 127, 137, 211 f
covariant, surface (a-metric), 13 ff, 19, 36, 44, 48, 52, 60, 71,
75, 98, 101 f, 113, 124 ff, 135 ff, 146, 154, 185 f, 203,
208 ff, 218 ff, 224, 225 f, 228 ff, 233
covariant, off surface (g-metric), 7, 71, 75, 92 ff, 101 f, 113,
127 ff, 137, 148, 203 f, 208 f, 213 ff, 218 ff, 222 ff, 228 ff, 233
Momentum equation, see also Navier-Stokes equations and Boundary-
layer equations, 71 f, 148, 151 f, 173
Momentum flux, 23, 75
Momentum-loss thickness, see Boundary-layer integral parameter
Navier-Stokes equations, 1 ff, 8 ff, 80 ff, 100, 102, 116, 146 ff,
231
thin-layer approximation, 5, 8, 80 ff, 91 ff, 203
Newtonian fluid, 11, 80
Noise, 3

- 262 -
Non-dimensionalization, 13 ff, 19, 118, 151, 199,214,226,228 ff
Normalizing length, 13, 119 ff, 131 ff
Nose radius, 107 ff, 179
No-slip condition , 1, 17, 20, 38, 146
Numerical solution, 1 f, 79, 62 ff, 88 ff, 116, 180 ff

Order of magnitude, 8, 70, 101

Parallel isobar design (of wing), 42, 52


Plane of symmetry (flow), 27, 34 ff, 56, 66 ff, 139, 157, 164 f,
177, 193 ff
Position vector, 199 ff
Prandtl number, 15, 20
Pressure, static, 14, 27 ff, 91 f, 102, 141, 231
gradient, 29, 38, 47, 70, 89, 149, 174, 180
relative-maximum line, 173 ff
Quantity
contravariant, 13, 228 ff
dimensionless, 228 ff
physical, 13, 228 ff
stretched, 228 ff
Ratio of specific heats, 15
Recirculatory flow, 116
Reference quantity, 13 f, 189, 197,228 ff
density, 15, 231
heat conductivity, 15, 232
length, 14, 118, 125, 214, 231 ff
pressure, 15,92,231
specific heats, 15, 231 f
speed of sound, 15
temperature, 15, 231
velocity, 13, 231
viscosity, 15, 232
Reynolds number, 1 ff, 8,13 ff, 86 f, 98, 111,182 ff, 214,224,229
Reynolds stress, 88, 91
Root of wing, see Wing
Rotational flow, 47, 52, 79

Second viscosity coefficient, 81, 149


Separation, 1 f, 9, 86, 100, 114, 141, 156 ff
criterion, 169
topology, 9, 156 ff
Separation bubble, 56, 114
Separation line, 156 ff, 163, 169, 176
angle, 160 ff
curvature, 164 ff
singular, 156
Separation point, singular, 156, 160 f, 164
Series expansion, 6b, 146, 221 ff
Shear stress, also wall shear stress, 2, 19 f, 30, 39, 46, 50, 54,
59, 73, 150, 156 f, 159 ff, 165, 175, 179
Newtons law, 19
non-dimensionalization, 19
physical, 19
line of relative ITI-minimum, 164 ff, 169 ff, 176, 179
tensor, 19, 150
SiJiftcr, 7, 71 ff, 218 ff

- 263 -
Similarity transformation, 9, 180 ff
Skewing of coordinates, 6, 73, 202, 205, 216
Skin-friction line, also limiting wall streamline, 63, 65, 156 ff,
161 ff, 236
curvature, 167
points-of-inflexion line, 168 ff
Solution
steady state, 89
transient unsteady, 89
Span of wing, see Wing
Specific heat, 14 f, 231 f
Speed of sound, 15
Stagnation line, see Attachment line
Stagnation point (nodal), also point of attachment, 7, 26 ff, 39, 42,
65 ff, 114, 138 ff, 163, 175, 190 ff, 193, 222 f
axisymmetric, 30
saddle, 26, 39
Step
forward facing, 116
backward facing, 116
Stream function, 237
Streamline, 9, 236 ff
curvature, 26 f
direction, 63, 65, 180
direction angle, 233
elevation angle, 160 ff, 238 f
external inviscid, 26, 63, 64 f, 157, 163, 172 ff
length element, 236
points-of-inflexion line, 175
projection on surface, 64, 236 f
Stream surface, 27 f, 161 f
Stress tensor, 80 ff, 147, 150
Suction, see Boundary-layer suction
Surface, 2, 4 ff, 38, 61, 71 ff, 88, 97,146 ff, 159 ff, 200ff, 216,
223, 229, 236 f
concave, 4 ff, 201 f
convex, 4 ff, 39, 201
cooling, 3,18,26,28
curvature, 1, 9, 20, 100 ff
element, 239
heating, 3, 18, 26, 28
hydraulically smooth, 20
impermeable, 20, 22, 146
no-slip condition, 1, 17, 20, 38, 146
Sweep angle, see Wing
Symmetrical flow, see also Plane of symmetry, 27, 36, 57, 60 f,
66 f, 110
Symmetry of variables, 36, 40
even, 3&
odd, 36

Temperature, static, 14, 17 f, 27 f, 38, 45,149,231


gradient, 17, 38,45, 149
Thermally perfect gas, 12, 80, 92, 149
Time, 14, 232
'rime-averaged flow variables, 3, 12, 89
Time-dependent equation, 3, 89
Tip of wing, see Wing
Transformation, 4, 33, 71 ff, 105, 122, 205 ff, 219 f, 229 f

- 264 -
Transport properties and coefficients, 3, 12, 93, 102, 149, 232
effective, 12, 149, 232
mULecular (laminar flow), 88
turbulent, 12, 102, 116, 149, 159
Turbomachine, 2
Turbulence
large eddy simulation, 88
model, 2 f, 20, 70, 97, 100, 112, 114
Turbulent flow, see Flow (boundary layer)
Turbulent Prandtl number, 20
Turbulent reattachment after separation, 114
Unit vector, 199 ff
Variables
dependent, 21, 29, 32, 36, 91, 146, 180, 191 ff, 228, 231, 237
independent, 2, 91, 228, 233 ff
primitive, 150
Vector coordinates (components)
contravariant, 3, 9, 13, 205 f, 221, 225, 228, 237
covariant, 13, 237
Velocity
chordwise, 56, 67
covariant coordinates (components), 13, 225, 237
contravariant coordinates (components) 3, 9, 13, 22, 33, 71.ff,
82, 91, 205 f, 223, 225, 229, 231, 237 f
distribution, 65
gradient, 28 f, 37 f, 110, 150, 191 f
physical, 13, 229, 231
profile, 2, 97 f, 180
spanwise coordinate (component), 42, 50, 52, 69
vector, 73 f, 205 f, 218 ff, 222 f, 236
Viscosity, 14, 232
apparent turbulent, 20
bulk, see Bulk viscosity
effective, 20, 232
molecular, 20
Viscous-inviscid interaction, 1 f, 23, 74, 98, 100, 116
Viscous sub-layer, 181
Volume element, 239
Vortex flow, 1,69,86,116,156
Vorticity vector, 98, 225 ff, 231
Wake flow, 1, 100
Wake region in turbulent boundary layer, 181
Wall-heat flux, see Heat flux
Wall temperature, 17, 45
Weingarten formula, 212, 220
Wing, 1, 5 f, 62, 66 ff, 96, 100, 103 ff, 110 ff, 116, 117 ff,
133, 157, 172 f
chord, 36, 48, 52, 56, 69, 111, 118
cross section, profile, 35, 119
finite, 34, 42 f, 52, 56, 67 f, 189
leading edge, 34 ff, 42, 44, 46, 48, 52, 67 f, 96, 103 ff, 107,
110 f, 114 f, 118 ff, 138, 157, 172 f, 176, 179, 189
lifting, 34, 56, 65
lower surface, 34, 48, 67, 121
non-lifting, 66

- 265 -
root, 42, 69, 103, 116, 118 ff
swept, 68, 103 ff, 111, 114, 122 ff, 170
sweep angle, 42, 44, 48, 52, 124, 176
tapered, 43, 52, 56, 114
tip, 42, 62, 65, 69, 103, 120
tip vortex, 69
trailing edge, 1, 5, 42, 108, 114, 118 ff, 179
upper surface, 34, 48, 67, 120 f
Wing-body combination, 2, 65
Wing-body junction, G, 62, 96, 116

- 206 -

Das könnte Ihnen auch gefallen